Worm domains and Fefferman space–time singularities

Worm domains and Fefferman space–time singularities

Journal of Geometry and Physics 120 (2017) 142–168 Contents lists available at ScienceDirect Journal of Geometry and Physics journal homepage: www.e...

654KB Sizes 0 Downloads 30 Views

Journal of Geometry and Physics 120 (2017) 142–168

Contents lists available at ScienceDirect

Journal of Geometry and Physics journal homepage: www.elsevier.com/locate/geomphys

Worm domains and Fefferman space–time singularities Elisabetta Barletta a , Sorin Dragomir a, *, Marco M. Peloso b a

Università degli Studi della Basilicata, Dipartimento di Matematica, Informatica ed Economia, Via dell’Ateneo Lucano 10, 85100 Potenza, Italy b Dipartimento di Matematica, Università degli Studi di Milano, Via C. Saldini 50, 20133 Milano, Italy

article

info

Article history: Received 29 March 2017 Received in revised form 5 June 2017 Accepted 6 June 2017 Available online 19 June 2017 Keywords: Worm domain Levi form Fefferman’s metric Curvature singularity Schmidt metric Bundle boundary

a b s t r a c t Let W be a smoothly bounded worm domain in C2 and let A = Null(Lθ ) be the set of Leviflat points on the boundary ∂ W of W . We study the relationship between pseudohermitian geometry of the strictly pseudoconvex locus M = ∂ W \ A and the theory of space–time singularities associated to the Fefferman metric Fθ on the total space of the canonical π circle bundle S 1 → C (M) −→ M. Given any point (0, w0 ) ∈ A, we show that every lift Γ (ϕ ) ∈ C (M), 0 ≤ ϕ − log |w0 |2 < π /2, of the circle Γw0 : r = 2 cos[log |w0 |2 − ϕ] in M, runs into a curvature singularity of Fefferman’s space–time (C (M), Fθ ). We show that Σ = π −1 (Γw0 ) is a Lorentzian real surface in (C (M), Fθ ) such that the immersion ι : Σ ↪→ C (M) has a flat normal connection. Consequently, there is a natural isometric immersion j : O(Σ ) → O(C (M), Σ ) between the total spaces of the principal bundles of Lorentzian frames O(1, 1) → O(Σ ) → Σ and adapted Lorentzian frames O(1, 1) × O(2) → O(C (M), Σ ) → Σ , endowed with Schmidt metrics, descending to a map of bundle completions which maps the b-boundary of Σ into the adapted bundle boundary of C (M), ˙ ) ⊂ ∂adt C (M). i.e. j(Σ © 2017 Elsevier B.V. All rights reserved.

1. Worm domains and Fefferman’s metric

{

}

A worm domain is a smoothly bounded pseudoconvex domain W in C2 given by W = (z , w ) ∈ C2 : ρ (z , w ) < 0 where

⏐ ( ) 2 ⏐2 ρ (z , w) = ⏐z − ei log |w| ⏐ − 1 + η log |w|2 ,

(1)

and η ∈ C (R) satisfies the conditions: (i) η(t) ≥ 0 and η is even and convex; (ii) η (0) = Iµ ≡ [−µ, µ] ⊂ R; (iii) there is a > 0 such that |t | > a yields η(t) > 1; and (iv) η(t) = 1 implies η′ (t) ̸ = 0. Originally devised by K. Diederich and J.E. Fornaess [1] to produce examples of pseudoconvex domains without a Stein neighborhood basis, worm domains are known (cf. e.g. [2]) to exhibit an array of peculiar irregularity properties. Perhaps the most remarkable of such properties, is the failure of satisfying Condition R, that is the condition that the Bergman projection −1



Pf (z , w ) =



K (z , w, ζ , ω) f (ζ , ω) dV (ζ , ω),

(z , w ) ∈ W ,

W

maps C ∞ W to C ∞ W (cf. M. Christ, [3]). Here K (z , w, ζ , ω) is the Bergman kernel of W and dV denotes the Lebesgue measure in C2 . It is well known that any smoothly bounded strictly pseudoconvex domain satisfies Condition R (see [4]

( )

*

( )

Corresponding author. E-mail addresses: [email protected] (E. Barletta), [email protected] (S. Dragomir), [email protected] (M.M. Peloso).

http://dx.doi.org/10.1016/j.geomphys.2017.06.001 0393-0440/© 2017 Elsevier B.V. All rights reserved.

E. Barletta et al. / Journal of Geometry and Physics 120 (2017) 142–168

143

and [5]) and indeed W is only weakly pseudoconvex, for the Levi form of its boundary vanishes at each point of the annulus A = {(0, w ) ∈ ∂ W : ⏐log |w|2 ⏐ ≤ µ}





(cf. Proposition 1.3 in [2], p. 480). The goal of this paper is to understand the nature of weakly pseudoconvex points on ∂ W by investigating the behavior of geometric objects as points approach the critical annulus A. We do this by combining methods in pseudohermitian geometry, as created by S.M. Webster [6], and the theory of space–time singularities, see. e.g. C.J.S. Clarke, [7], by profiting from the presence of the Fefferman metric of W , a Lorentzian metric defined solely on a saturated open subset of the total space of the canonical circle bundle S 1 → C (∂ W ) → ∂ W . If U(z , w, ζ ) = |ζ |2/(n+1) u(z , w ),

u(z , w ) = K (z , w, z , w )−1/(n+1) ,

(2)

with n = 2 then one may equip W × (C \ {0}) with the (0, 2)-tensor field G=

2 ∑

∂ 2U

∂ zA ∂ zB A,B=0

dz A ⊙ dz ,

z0 = ζ ,

B

z1 = z,

z 2 = w,

(3)

as devised by C. Fefferman (cf. [8]) by following the ideas of I. Naruki [9] as to associating to W the suspended domain

˜ = {(z , w, ζ ) ∈ W × C : |ζ |2(n+1) K (z , w, z , w) < 1}. W The pullback of G to W × S 1 , that is, j∗ G where j : W × S 1 ↪→ W × (C \{0}), is certainly degenerate. A fundamental question is then the following: Does j∗ G approach a nondegenerate (perhaps Lorentzian) metric F on ∂ W × S 1 as (z , w ) → ∂ W ? The answer in [8], that F is actually a Lorentzian metric on ∂ Ω × S 1 , is provided only for the case of a smoothly bounded strictly pseudoconvex domain Ω ⊂ C2 and relies on the asymptotic expansion of the Bergman kernel [10]. For ϵ > 0, the approximating domains Wϵ = {(z , w ) ∈ C2 : ρ (z , w ) + ϵ < 0}

are strongly pseudoconvex, as we will easily observe, and Fefferman’s asymptotic expansion formula applies to the Bergman kernel Kϵ (z , w, ζ , ω) to give Kϵ (z , w, z , w ) = cWϵ |∇ρ (z , w )|2 · det Lϵ (z , w ) · |ρϵ (z , w )|−3 + Eϵ (z , w, z , w )

(4)

for some constant cWϵ > 0 and some function Eϵ (z , w, ζ , ω) such that Eϵ ∈ C (W ϵ × W ϵ \ ∆ϵ ) and ∞

⏐ ⏐ ⏐ ⏐ ⏐Eϵ (z , w, z , w)⏐ ≤ c ′ |ρϵ (z , w)|−3+ 12 · ⏐ log|ρϵ (z , w)|⏐. Wϵ ′ Here Lϵ is the restriction of ∂∂ρ to T1,0 (∂ Wϵ ) ⊗ T0,1 (∂ Wϵ ), ∆ϵ is the diagonal of ∂ Wϵ × ∂ Wϵ , and cW > 0 is another constant ϵ (depending on Wϵ ). As an elementary consequence of the expansion (4),

ϕϵ (z , w) = −Kϵ (z , w, z , w)−1/3 is a defining function for Wϵ , though solely of class C 2 , implying that Fϵ =

lim

(z ,w )→∂ Wϵ

j∗ϵ Gϵ

is a Lorentzian metric on ∂ Wϵ × S 1 , whose restricted conformal class is a biholomorphic invariant of Wϵ , cf. [8]. It is of great interest to obtain some analog to the expansion (4) in the case of smooth, bounded, weakly pseudoconvex domains, and in particular for the worm domain, corresponding to the case ϵ = 0. An alternative approach is provided in [8] as well and relies on the possibility of replacing u(z , w ) in (2)–(3) by the (eventual) solution to the Dirichlet problem for the complex Monge–Ampère equation,

{

J(u) = 1 in W u = 0 on ∂ W ,

(5)

where u J(u) = (−1) det uz uw

(

n

uz uzz uw z

uw uz w uww

) .

The solution by S.-Y. Cheng & S.-T. Yau [11] to the problem (5) is confined to the case of a strictly pseudoconvex domain. S.-Y. Li [12] studied the same problem in the weakly pseudoconvex case, yet his approach requires a defining function which is plurisubharmonic on the entire boundary of the given pseudoconvex domain and none exist for W , cf. Proposition 2.2 in [2], p. 486. The solutions in [11] and [12] were unknown at the time [8] was published yet, as observed in [8], only the

144

E. Barletta et al. / Journal of Geometry and Physics 120 (2017) 142–168

two-jet at the boundary of a solution u(z , w ) to (5) is required in the construction of F and an approximate solution of the sort may be built by setting

{ u=

1+

1 4

[

(

1−J

ρ J(ρ )

)]}

ρ J(ρ )

(6)

so that u = 1 + O(ρ 2 ), provided that (6), and in general the approximation scheme in [8], make sense for Ω = W . An elementary calculation shows that J(ρϵ )(0,w0 ) = −

ϵ |w0 |

for any ϵ ≥ 0 and any |log |w0 |2 | ≤ µ. In particular J(ρ ) vanishes along the annulus A and the alternative choice (6) is not available. Thus, in the context of the worm domain W ⊂ C2 , one misses the tools for understanding the behavior of j∗ G in the limit as (z , w ) → ∂ W and the extrinsic approach (as outlined above, cf. [8]) to a Fefferman-like metric on ∂ W × S 1 is an open problem. In this paper we consider M = ∂ W \ A, which an open subset of ∂ W and turns out to be a strictly pseudoconvex CR manifold. Observing that 2

2

ρ (z , w) = η(log |w|2 ) − e−i log |w| z − ei log |w| z + zz , we see that

θ=

i 2

j∗ ∂ − ∂ ρ ,

(

)

j : ∂ W ↪→ C2 ,

(7)

is a pseudohermitian structure on ∂ W whose pointwise restriction to M is a positively oriented contact form on M. Fefferman’s question (cf. [8]) whether an intrinsic approach to a Lorentzian metric, perhaps living on M × S 1 , is available when M is not the boundary of a domain was settled by J.M. Lee (cf. [13]). π Consider the total space C (M) of the canonical circle bundle S 1 → C (M) → M, and indeed it is globally diffeomorphic 1 to M × S . In [13] Lee built a Lorentzian metric Fθ ∈ Lor[C (M)] in terms of pseudohermitian invariants of (M , θ ), which is conformal equivalent to the (extrinsic) Fefferman metric, when the latter is defined. The Lorentzian manifold (C (M), Fθ ) admits a natural time orientation, hence it is a space–time. We shall provide details of this construction in Section 3. Let Γ : [a, b) → C (M) be a timelike curve in (C (M), Fθ ), parametrized by proper time t. Following (cf. [14]), we say that Γ runs into a curvature singularity if (i) Γ has bounded acceleration, (ii) λ ◦ Γ˙ is unbounded for some curvature invariant λ. The physical interpretation is that Γ , thought of as the world line for some particle or moving observer, encounters unbounded curvature as t → b and hence cannot be extended to proper time t = b, not in the space–time C (M) or any of its smooth extensions. If γ : [a, b) → M is a C 1 curve, by a lift of γ one means a C 1 curve Γ : [a, b) → C (M) such that π (Γ (t)) = γ (t) for any a ≤ t < b. ⏐ ⏐ Let w0 ∈ C \ {0} such that ⏐log |w0 |2 ⏐ ≤ µ. Let us set

[

I(w0 ) = log |w0 |2 , log |w0 |2 +

π) 2

⊂R

and let us consider

( [ ] ) γw0 (ϕ ) = 2 cos ϕ − log |w0 |2 eiϕ , w0 ,

ϕ ∈ I(w0 ),

which is a circle in M such as limϕ→log |w0 |2 +π /2 γw0 (ϕ ) does not exist in M — yet the limit is (0, w0 ) ∈ A in the topology of

∂W.

Finally, let D be the Levi-Civita connection of the Lorentzian manifold (C (M), Fθ ) and K its scalar curvature. Our first main result is the following. Theorem 1. Every lift Γ (ϕ ) ∈ C (M) of γw0 (ϕ ) ∈ M runs into a curvature singularity. Precisely, lim

ϕ→log |w0 |2 +π /2

K (Γ (ϕ )) = −∞,

(8)

i.e. Γ (ϕ ) encounters unbounded scalar curvature as ϕ → log |w0 |2 + π/2. Theorem 1 is proved in Section 3. Notice that Theorem 1 holds for an arbitrary lift of the circle γw0 . We may nevertheless show that γw0 may be lifted to a timelike curve Γ in (C (M), Fθ ), thus relating Theorem 1 to the general relativistic notions in [14].

E. Barletta et al. / Journal of Geometry and Physics 120 (2017) 142–168

145

Theorem 2. There exist C ∞ curves Γ : I(w0 ) → C (M) such that: (i) π (Γ (ϕ )) = γw0 (ϕ ) for any ϕ( ∈ I(w0 ); ) (ii) Γ is timelike, that is, Fθ ,Γ (ϕ ) Γ˙ (ϕ ) (, Γ˙ (ϕ ) < 0 for) every ϕ ; (iii) Γ is future directed, that is, Fθ ,Γ (ϕ ) Γ˙ (ϕ ) , Xθ , Γ (ϕ ) < 0 for every ϕ , where Xθ is a time orientation of (C (M), Fθ ). The construction of the canonical time orientation Xθ of (C (M), Fθ ) is also deferred to Section 3. Theorem 2 is proved in Section 4. According to J.A. Thorpe points where a timelike or null curve Γ of bounded acceleration runs into a curvature singularity are b-boundary points. No proof of the statement is provided in [14]. Nevertheless, in view of Theorem 1, one may expect that limϕ→log |w0 |2 +π /2 Γ (ϕ ) exists in the topology of the b-completion C (M) and lies on the b-boundary C˙ (M) in the sense of G.B. Schmidt, [15]. The opposite scenario is however suggested by the following Theorem 3. The circle γw0 : I(w0 ) → M has infinite energy

( ) E γw0 =

log |w0 |2 + π2



log |w0

|2

gθ , γ (ϕ ) γw ˙ 0 (ϕ ) , γw˙ 0 (ϕ ) dϕ = +∞

(

)

with respect to the Webster metric gθ on M. Therefore an observer attached to γw0 requires infinite energy to reach (0, w0 ) ∈ A. Theorem 3 is proved in Section 4. From now on, we write M = (C (M), Fθ ) for short, and let O(1, 3) → O(M) → M be the principal bundle of all Lorentzian frames. Let O++ (1, 3) and O+ (M) be connected components of O(1, 3) and O(M) respectively. Following [15], we let dG : O+ (M) × O+ (M) → [0, +∞) denote the Schmidt distance function and O+ (M) the ˙ of M is given by Cauchy completion of O+ (M) with respect to dG . The b-boundary M

]

[

˙ = O+ (M)/O++ (1, 3) \ M. M By a result in [15], b-boundary points may be characterized as end points of inextensible curves Γ : [a, b) → M, that is curves Γ for which limt →b− Γ (t) does not exist in the topology of M, whose horizontal lifts Γ H : [a, b) → O+ (M) have finite length b



GΓ H (t) Γ˙ H (t) , Γ˙ H (t)

(

a

)1/2

dt < ∞

(9)

with respect to the Schmidt metric G (cf. [15]), which is a Riemannian metric on O+ (M) determined by the Levi-Civita connection 1-form of the Fefferman metric Fθ . Let Z = −ρw ∂/∂ z + ρz ∂/∂w be a generator of the CR structure T1,0 (∂ W ) and let Z = X − iY be its real and imaginary parts. Also let us set S = [(n + 2)/2] ∂/∂ s

(10)

with n = 1 (the convention in [16], p. 083504-19). Let Σ : M → O (M) be the cross section given by +

Σ (p)eα = Eα (p) ∈ Tp{(M), p ∈ M, } {Eα : 0 ≤ α ≤ 3} = T ↑ − S , E1↑ , E2↑ , T ↑ + S , √ √ 2

E1 =

g11

X,

E2 =

2

g11

(11)

Y.





Indeed, T ↑ − S is timelike while E1 , E2 and T ↑ + S are spacelike hence {Eα : 0 ≤ α ≤ 3} is a field of Lorentzian frames. Also, T ↑ and S are null. We are unable to prove or disprove (9) when Γ is a lift of γ = γw0 . However, the natural lift C = Σ ◦ Γ of any such Γ may be shown to have infinite length. Theorem 4. Let w0 ∈ C \ {0} such that ⏐log |w0 |2 ⏐ ≤ µ and let Γ : I(w0 ) → M be an arbitrary lift of the circle γw0 . If C : I(w0 ) → O+ (M) is defined by C = Σ ◦ Γ then





log |w0 |2 + π2 log |w0

|2

GC (ϕ ) C˙ (ϕ ) , C˙ (ϕ )

(

)1/2



dϕ = +∞

where G is the Schmidt metric on O+ (M). Theorem 4 is proved in Section 5. The question whether Π −1 (A) contains b-boundary points of Fefferman’s space–time ˙ ̸= ∅] remains open. [i.e. whether Π −1 (A) ∩ M

146

E. Barletta et al. / Journal of Geometry and Physics 120 (2017) 142–168

So far we have exhibited (cf. Theorem 1) only curvature singularities of the Fefferman space–time M. Section 6 demonstrates a dimension reduction argument (similar in spirit to the approach in [17] and [18]) leading to the identification of subsets of the adapted bundle boundary ∂adt M. Theorem 5. Let Σ ⊂ M be a Lorentzian surface with a flat normal connection (R⊥ = 0). Let O(1, 1) → O(Σ ) → Σ and O(1, 1) × O(2) → O(M, Σ ) → Σ be respectively the principal bundle of Lorentzian frames tangent to Σ and the principal bundle of adapted Lorentzian frames tangent to M. Given a globally defined cross section sT : Σ → O(Σ ) there is a natural isometric immersion

j : O(Σ ) → O(M, Σ ),

j∗ Gadt = GΣ ,

(12)

with respect to the Schmidt metrics GΣ and Gadt . j descends to a map of orbit spaces

j:Σ =

O+ (Σ ) O++ (1, 1)



O+ (M, Σ ) O++ (1, 1)

× O+ (2)

=M

adt

(13)

such that

˙ ) ⊂ ∂adt M j(Σ

(14)

˙ = Σ \ Σ and ∂adt M = M adt \ M. where Σ The construction of the Schmidt metric Gadt on O(M, Σ ) and of the corresponding distance function dadt on O+ (M, Σ ) relies on the connection 1-form ω obtained as the o(1, 1) ⊕ o(2) component of i∗ j∗ ψ where ψ is the Levi-Civita connection 1-form of the Fefferman space–time M while j

i

O(M, Σ ) −→ O(M)|Σ −→ O(M) are injections. This is because i∗ j∗ ψ is not a connection 1-form on O(M, Σ ) and in particular the resulting bundle completions ˙ and ∂adt M) are logically unrelated. Theorem 5 is proved in Section 6. and bundle boundaries (i.e. M Section 7 is devoted to the geometry of the total space Σ of a principal circle subbundle S 1 → Σ → Γw , in S 1 → M → M, over the circle

Γw = {γw (ϕ ) : ϕ ∈ I(w ) \ {log |w|2 }},

⏐ ⏐ ⏐log |w|2 ⏐ < µ.

Such real surfaces Σ are shown to be examples to which our Theorem 5 applies. Theorem 6. For every |f (w )| < µ the immersion

ι : π −1 (Γw ) ↪→ M = (C (M), Fθ ) has a Lorentzian first fundamental form and a flat normal connection. Let sT : Σ → O(Σ ) be the section defined by sT (p)e0 = (dγ ↑ (ϕ ) Rζ )γ˙w↑ (ϕ ), w

p = γw↑ (ϕ ; p0 · ζ ) ∈ Σ ,

sT (p)e1 = Sp ,

p0 ∈ Σ ,

ζ ∈ S1 ,

2 4 where {e0( , e1 } ⊂ R orthonormal frame in the normal )⊥×{0} ⊂ R is⊥ the canonical linear basis. Let {N1 , N2 } be a globally defined −1 bundle T π (Γw ) such that ∇ Ni = 0 for i ∈ {1, 2} and let us consider the section s⊥ : Σ → O(T (Σ )⊥ ) defined by

s⊥ (p) : R2 → T π −1 (Γw )

(

s (p)ei+1 = Ni (p),

)⊥ p

p∈π



,

−1

(Γw ), i ∈ {1, 2},

where {e2 , e3 } ⊂ {0} × R ⊂ R . The sections (sT , s⊥ ) determine an isometric immersion 2

4

j : O π −1 (Γw ) → O M, π −1 (Γw )

(

)

(

)

descending to a map of bundle completions

j : π −1 (Γw ) → M

adt

which maps the b-boundary points of (π −1 (Γw ), , ι∗ Fθ ) into the adapted bundle boundary ∂adt M. The property j∗ Gadt = GΣ , where GΣ and Gadt are the Schmidt metrics (cf. [15]) on O(Σ ) and O(M, Σ ), claimed in Theorem 6 is essentially a consequence of

( ⊥ )∗ s

ψ⊥ = 0

(15)

where ψ is the normal connection 1-form of the immersion Σ ↪→ M. In turn (15) follows from the fact that the frame {N1 , N2 } determining s⊥ is parallel in the normal bundle. Such a frame may be chosen provided that Σ ↪→ M has a flat normal connection, which is the case for Σ = π −1 (Γw ) according to our Lemma 14. ⊥

E. Barletta et al. / Journal of Geometry and Physics 120 (2017) 142–168

147

2. CR and pseudohermitian geometry We shall prove Theorems 1 and 2 by using techniques in pseudohermitian geometry (cf. e.g. [19]). We start by describing the CR and pseudohermitian structures on ∂ W and M. The CR structure T1,0 (∂ W )(z ,w) = T(z ,w) (∂ W )⊗R C ∩ T 1,0 (C2 )(z ,w) ,

[

(z , w ) ∈ ∂ W ,

]

2

induced by the complex structure on C , is globally the span of Z = −ρw ∂/∂ z + ρz ∂/∂w

(16)

where ρ is given by (1) and ρz = ∂ρ/∂ z etc. Let H(∂ W ) = Re T1,0 (∂ W ) ⊕ T0,1 (∂ W ) ,

{

T0,1 (∂ W ) ≡ T1,0 (∂ W ) ,

}

be the Levi, or maximally complex, distribution on ∂ W , thought of as endowed with the complex structure J : H(∂ W ) → H(∂ W ),

J V +V =i V −V ,

)

(

(

)

V ∈ T1,0 (∂ W ).

Let E(∂ W ) ⊂ T (∂ W ) be the conormal bundle associated to the Levi distribution i.e. the real line bundle given by ∗

E(∂ W )(z ,w) = ω ∈ T(z∗ ,w) (∂ W ) : Ker(ω) ⊃ H(∂ W )(z ,w)

{

}

for any (z , w ) ∈ ∂ W . Then θ , defined in (7), is a pseudohermitian structure on ∂ W i.e. a globally defined nowhere zero C ∞ section in E(∂ W ). The Levi form is Lθ V , W = −i (dθ ) V , W

(

)

(

)

for any V , W ∈ T1,0 (∂ W ). We also set Gθ (X , Y ) = (dθ )(X , JY ) for any X , Y ∈ H(∂ W ), so that Lθ is the same as the C-linear extension of Gθ to T1,0 (∂ W ) ⊗ T0,1 (∂ W ). Lemma 1. Let us set f (w ) = log |w|2 for any w ∈ C \ {0}. (i) The Levi invariant g11 = Lθ (Z , Z ) ∈ C ∞ (∂ W ) may be expressed as the third degree polynomial in z and z Fα1 α 2 (w ) z α1 z



g11 =

α2

(17)

0≤|α|≤3

with Hermitian coefficients Fmn ∈ C ∞ (C, C) i.e. Fmn = Fnm given by F00 (w ) =

1

F20 (w ) = −

F21 (w ) =

[

2

F10 (w ) = −

F11 (w ) =

e−f (w) η′′ (f (w )) + η′ (f (w ))2 , 1 2 1 2

1 2 1

2 F30 (w ) = 0.

]

e−(1+i)f (w) η′′ (f (w )) − i η′ (f (w )) − 1 ,

[

]

e−(1+2i)f (w) ,

(18)

[ ] e−f (w) η′′ (f (w )) − 2 , e−(1+i)f (w) ,

(ii) If A = {(z , w ) ∈ ∂ W : (Lθ )(z ,w) = 0} is the null space of the Levi form, then A = {(0, w ) ∈ ∂ W :⏐ |f (w )| ≤ µ}. (iii) M = ∂ W \ A is a strictly pseudoconvex CR manifold, with the CR structure T1,0 (M) = T1,0 (∂ W )⏐M . (iv) Let Wϵ = {z , w ) ∈ C2 : ρ (z , w ) < −ϵ} with ϵ ≥ 0. There is ϵ0 > 0 such that {Wϵ : 0 ≤ ϵ ≤ ϵ0 } is a family of pseudoconvex domains with Wϵ strictly pseudoconvex for every 0 < ϵ ≤ ϵ0 . Proof. We first observe that the identity i {

(dθ )(Z , Z ) =

2

|ρw |2 ρzz + |ρz |2 ρww − ρw ρz ρz w − ρz ρw ρwz

}

together with

ρz = z − e−if (w) ,

ρzz = 1,

ρz w = i fw e−i f (w) ,

yields g11 =

i {( 2

+

1 − z e−i f (w) fw ρw − 1 − z ei f (w) fw ρw

)

1{ 2

(

)

}

} |ρw |2 + ρww (1 + |z |2 − z e−if (w) − z eif (w) ) .

(19)

148

E. Barletta et al. / Journal of Geometry and Physics 120 (2017) 142–168

Since

{ } ρw = fw η′ (f (w)) + i(z e−if (w) − z eif (w) ) , and

{ } ρww = |fw |2 η′′ (f (w)) + z e−i f (w) + z ei f (w) , Eq. (17) follows. Next for every ϵ ≥ 0 let ϕϵ be the local defining function for Wϵ given by

ϕϵ (z , w) = earg(w ) ρϵ (z , w) 2

with ρϵ = ρ + ϵ , so that ϕϵ = ϕ (1) + ϕ (2) + ϕϵ(3) where

ϕ (1) (z , w) = |z |2 earg(w ) , 2

ϕ (2) (z , w) = −2 Re(z e−i logw ), 2

ϕϵ(3) (z , w) = [η(f (w)) + ϵ ] earg(w ) . 2

The (locally well defined) function e− log w is holomorphic and its modulus is earg(w ) hence ϕ (1) and ϕ (2) are plurisubharmonic. Also ∆ϕϵ(3) ≥ 0 so ϕϵ(3) is plurisubharmonic, as well. So for ϵ0 > 0 sufficiently small each Wϵ , 0 ≤ ϵ ≤ ϵ0 , is a pseudoconvex domain, as claimed. Actually if ϵ > 0 then ϕϵ is strictly plurisubharmonic at every point (z , w ) because of ϕϵ(3) , while if ϵ = 0 then ϕ0 is strictly plurisubharmonic at each point (z , w) with z ̸= 0 (because of ϕ (1) and ϕ (2) ) or with |f (w)| > µ (because of ϕ0(3) ). Hence each Wϵ , 0 < ϵ ≤ ϵ0 is a strictly pseudoconvex domain and the weak pseudoconvexity locus of W is contained in 2

2

A = {(0, w ) : ⏐log |w|2 ⏐ ≤ µ} ⊂ ∂ W .





These are precisely the points where the Levi form Lθ vanishes, may be seen as a consequence of (17) and (18). This proves Lemma 1. □ By Lemma 1 the pullback of θ to M (denoted by the same symbol θ ) is a contact form on M i.e. θ ∧ dθ is a volume form on M. Let T ∈ X(M) be the Reeb vector field i.e. the globally defined, nowhere zero, vector field on M, transverse to the Levi distribution H(M), determined by

θ (T ) = 1, T ⌋ dθ = 0. Then T (M) = H(M) ⊕ RT . Let gθ be the Webster metric i.e. the Riemannian metric on M given by gθ (X , Y ) = Gθ (X , Y ),

gθ (X , T ) = 0,

gθ (T , T ) = 1,

for any X , Y ∈ H(M). Let ∇ be the Tanaka–Webster connection of (M , θ ) i.e. the unique linear connection on M satisfying: (i) the Levi distribution H(M) is parallel with respect to ∇ [i.e. Y ∈ H(M) H⇒ ∇X Y ∈ H(M) for any X ∈ X(M)]; (ii) ∇ T = 0, ∇ gθ = 0; (iii) the torsion tensor field T∇ is pure i.e. for any V , W ∈ T1,0 (M) and X ∈ X(M), T∇ (V , W ) = 0,

T∇ V , W = 2i Lθ V , W ,

(

)

(

)

τ ◦ J + J ◦ τ = 0,

where τ (X ) ≡ T∇ (T , X ) is the pseudohermitian torsion of (M , θ ). Let R∇ be the curvature tensor field of ∇ , defined as R∇ (X , Y ) = [∇X , ∇Y ] − ∇[X ,Y ] for any X , Y ∈ X(M). The pointwise restriction of Z [given by (16)] to M gives a frame T1 in T1,0 (M) and one sets A ∇ TB = ωB A ⊗ TA , ∇TB TC = ΓBC TA ,

where A, B, C , . . . ∈ 0, 1, 1 ,

{

}

T0 = T ,

T1 = Z ,

so that ωB and are respectively the connection 1-forms and Christoffel symbols of the Tanaka–Webster connection ∇ . As ∇ parallelizes both H(M) and J, it parallelizes the eigen-distributions T1,0 (M) and T0,1 (M) hence ω1 1 = 0 and ω1 1 = 0. Also ∇ T = 0 yields ω0 A = 0. Hence all 1-forms ωA B vanish except for ω1 1 and their complex conjugates ω1 1 = ω1 1 . Let θ 1 be the complex 1-form on M determined by A

A ΓBC

θ 1 (Z ) = 1,

θ 1 (Z ) = 0,

θ 1 (T ) = 0,

E. Barletta et al. / Journal of Geometry and Physics 120 (2017) 142–168

149

and let us set θ 1 = θ 1 . Then {θ 1 , θ 1 , θ } is a global frame of T ∗ (M) ⊗ C and A θB , ωC A = ΓBC

θ 0 = θ.

As to the curvature components with respect to the frame {TA }, one sets R∇ T1 , T1 T1 = R1 1 11 T1 ,

(

)

R1111 = g11 R1 1 11 ,

Ric∇ (X , Y ) = Trace V ↦ → R∇ (V , Y )X ,

{

RAB = Ric∇ (TA , TB ),

}

R = g 11 R11 ,

X , Y ∈ X(M),

g 11 ≡ 1/g11 .

The terms R11 and R are respectively the pseudohermitian Ricci tensor and the pseudohermitian scalar curvature of (M , θ ). By a result of S.M. Webster (cf. [6]) τ maps T1,0 (M) into T0,1 (M) hence one may set

τ (T1 ) = A1 1 T1 ,

A11 = g11 A1 1 .

Lemma 2. The Christoffel symbols of the Tanaka–Webster connection of (M , θ ) are given by 1 = g 11 Z (g11 ), Γ11

1 Γ11 = 0,

(20)

1 1 2iρz g11 Γ01 = Γ11 (ρz ρz w − ρw )

( ) + ρz w (ρz ρz w − ρw ) − |ρz |2 ρz ww + 2 ρz ρww − |ρz w |2 .

(21)

Also Hα1 α 2 (w ) z α1 z



Z (g11 ) =

α2

(22)

0≤|α|≤3

where Hmn ∈ C ∞ (C, C) are given by 1

η′ (f (w))F10 (w) − e−f (w) G00 (w), w ] 1 [ H10 (w ) = −2η′ (f (w))F20 (w) − ie−if (w) F10 (w) − −e−if (w) G10 (w), w ] 1 [ ′ −η (f (w))F11 (w) + ieif (w) F10 (w) + G00 (w) − e−if G01 (w), H01 (w ) = w H00 (w ) = −

H20 (w ) = −

2i

H21 (w ) = −

2i

e−if (w) F20 (w ) − e−if (w) G20 (w ), w ] 1 [ −ie−if (w) F11 (w) + 2ieif (w) F20 (w) − 2η′ (f (w))F21 (w) H11 (w ) = w + G10 (w ) − e−if (w) G11 (w ), ] 1 [ ′ −η (f (w))F12 (w) + ieif (w) F11 (w) + G01 (w) − e−if (w) G02 (w), H02 (w ) = w H30 (w ) = 0,

H12 (w ) = −

i

w

e−if (w) F21 (w ) + G20 (w ) − e−if (w) G21 (w ), e−if (w) F12 (w ) + G11 (w ) − e−if (w) G12 (w ),

eif (w) F12 (w ) + G02 (w ), w = ∂ Fmn /∂w for any 0 ≤ m + n ≤ 3.

H03 (w ) = and Gmn

i

w

Proof. Let us consider the functions 1 ρw − ρz ρz w α= , β = H α, H = . ρz − H ρw ρw ρz w − ρz ρww The Reeb vector field of (M , θ ) is T = i α ∂z − α ∂z − β ∂w + β ∂w .

{

}

(23)

By the purity axiom (in the description of ∇ )

[ ] 1 1 Γ11 Z − Γ11 Z − 2ig11 T = Z , Z .

(24)

150

E. Barletta et al. / Journal of Geometry and Physics 120 (2017) 142–168

On one hand Z , Z = (ρz ρww − ρw ρz w ) ∂z + (ρw ρzz − ρz ρz w ) ∂w

[

]

(25)

+ (ρw ρz w − ρz ρww ) ∂z + (ρz ρz w − ρw ρzz ) ∂w and on the other, by (23),

( 1 ) ( 1 ) 1 1 Z − Γ11 Z − 2ig11 T = Γ11 ρw + 2 α g11 ∂z − Γ11 ρz + 2 β g11 ∂w Γ11 ( ( ) ) 1 1 − Γ11 ρw + 2 α g11 ∂z + Γ11 ρz + 2 β g11 ∂w .

(26)

Substitution from (25)–(26) into (24) leads to 1 ρz ρww − ρw ρz w = Γ11 ρw + 2 α g11 ,

(27)

1 ρw ρzz − ρz ρz w = −Γ11 ρz − 2 β g11 .

(28)

Summing up (27)–(28) furnishes g11 =

1{

|ρz |2 ρww + |ρw |2 − ρz ρw ρz w − ρz ρw ρz w

2

}

(29)

which is easily seen to agree with (19). Next we substitute from (29) into (27) and obtain Γ 1 = 0, which is the second 11 identity in (20). In particular (again by (24)) Z , Z = −2i g11 T

[

]

(30)

which is often taken as a definition for the Levi invariant g11 . By a result in [19]

{ } 1 Γ11 = g 11 Z (g11 ) − gθ (Z , [Z , Z ]) which together with (30) yields the first identity in (20). To prove (22) one starts from

∂ g11 = F10 (w) + 2F20 (w)z + F11 (w)z + 2F21 (w)zz + F12 (w)z 2 , ∂z ∑ ∂ g11 α = Gα1 α 2 (w ) z α1 z 2 , ∂w

(31)

0≤|α|≤3

and exploits (18) to obtain G00 (w ) =

1 2w

G10 (w ) = −

G01 (w ) = −

G20 (w ) = G11 (w ) =

e−f (w) η′′′ (f (w )) − η′′ (f (w )) + 2η′′ (f (w ))η′ (f (w ))

{

} − η′ (f (w))2 ,

1 2w 1 2w

1 2w 1 2w 1

e−(1+i)f (w) η′′′ (f (w )) − (1 + 2i)η′′ (f (w ))

{

} − (1 − i)η′ (f (w)) + 1 + i , e−(1−i)f (w) η′′′ (f (w )) − (1 − 2i)η′′ (f (w ))

{

} − (1 + i)η′ (f (w)) + 1 − i ,

(1 + 2i)e−(1+2i)f (w) , e−f (w) η′′′ (f (w )) − η′′ (f (w )) + 2 ,

{

}

(1 − 2i)e−(1−2i)f (w) , 2w 1 G21 (w ) = − (1 + i)e−(1+i)f (w) , 2w 1 G12 (w ) = − (1 − i)e−(1−i)f (w) , 2w G30 (w ) = 0, G03 (w ) = 0.

G02 (w ) =

We are left with the proof of (21). By the very definition of pseudohermitian torsion, 1 [T , Z ] = Γ01 Z − A1 1 Z .

(32)

E. Barletta et al. / Journal of Geometry and Physics 120 (2017) 142–168

151

On the other hand, by (23),

{ } [T , Z ] = i α ρz w − β ρww − α ρz w + β ρww + αz ρw − αw ρz ∂z { } + i α ρzz − β ρz w − α ρzz + β ρz w − βz ρw + βw ρz ∂w } } { { ∂α ∂β ∂β ∂α + ρz ∂z + i ρw − ρz ∂w + i −ρw ∂z ∂w ∂z ∂w

(33)

and 1 1 Z − A1 1 Z = Γ01 Γ01 (−ρw ∂z + ρz ∂w ) − A1 1 (−ρw ∂z + ρz ∂w ) .

(34)

Substituting (33) and (34) into (32), using the facts that ρzz = 0 and ρzz = 1, gives

{ } 1 Γ01 ρz = i −β ρz w − α + β ρz w + Z (β ) ,

(35)

and

} ∂β ∂β A1 ρz = i −ρw + ρz . ∂z ∂w {

1

Formulas (27), (28) with Γ 1 = 0 become 11

α=

1 2

g

11

{ρz ρww − ρw ρz w } ,

β=

1 2

g 11 {ρz ρz w − ρw ρzz } .

(36)

Moreover, by (36) and (20) it follows Z (β ) = −

+

1 2 1

1 g 11 Γ11 (ρz ρz w − ρw )

2

(37)

g 11 |ρz |2 ρz ww + ρz |ρz w |2 − ρww

{

(

)}

.

Finally, substituting (36) and (38) into (35) yields (21). □ 3. Pseudohermitian scalar curvature The curvature R1 1 11 and pseudohermitian Ricci curvature R11 of (M , θ ) are given by (cf. [6] or [19]) 1 1 1 1 1 1 1 ) + Γ11 ) − Z (Γ11 Γ11 − Γ11 Γ11 + 2i g11 Γ01 R11 = R1 1 11 = Z (Γ11 .

(38)

Consequently (by (20)) 1 1 ) + 2i g11 Γ01 R1 1 11 = −Z (Γ11 .

(39)

Let w0 ∈ C \ {0} such that log |w0 |2 ∈ Iµ . Let D(z0 ) = {z ∈ C : |z − z0 | < 1} be the unit disc of center z0 ∈ C and let Cw0 ⊂ ∂ W be the circle

[

(

Cw0 = ∂ D ei log |w0 |

=

{(

2

)]

× {w0 } [

reiϕ , w0 : r = 2 cos log |w0 |2 − ϕ , ⏐ϕ − log |w0 |2 ⏐ ≤

)

] ⏐



π} 2

.

Then Cw0 ∩ A = {(0, w0 )}. The key result, leading to (8) in Theorem 1, is the following. Proposition 1. Let w0 ∈ C \ {0} such that |f (w0 )| ≤ µ. Then R11 (reiϕ , w0 ) = −

2

|w0 |

+ O(r),

r → 0.

(40)

r → 0.

(41)

Consequently R(r eiϕ , w0 ) = −

4 r2

+ O(r −1 ),

Proof. The function η vanishes at f (w0 ) = log |w0 |2 to infinite order. Lemma 1 then gives 2 g11 (z , w0 ) = e−(1+i)f (w0 ) z + e−(1−i)f (w0 ) z − e−(1+2i)f (w0 ) z 2

− e−f (w0 ) zz − e−(1−2i)f (w0 ) z 2 + e−(1+i)f (w0 ) z 2 z + e−(1−i)f (w0 ) zz 2

152

E. Barletta et al. / Journal of Geometry and Physics 120 (2017) 142–168

for any z ∈ ∂ D(ei f (w0 ) ). The above equation in polar coordinates z = rei ϕ is 1

g11 (z , w0 ) =

2

r 2 e−f (w0 ) .

(42)

To compute Z (g11 )(z ,w0 ) one uses (22) and the equations F00 (w0 ) = 0,

F11 (w0 ) = −e−f (w0 ) , G00 (w0 ) = 0, G01 (w0 ) = − G11 (w0 ) =

1

F10 (w0 ) =

2

F21 (w0 ) =

G10 (w0 ) = − 1−i

1

1+i 2 w0

e−(1−i)f (w0 ) ,

2w0

w0

e−(1+i)f (w0 ) ,

e−f (w0 ) ,

2

e−(1+i)f (w0 ) ,

G02 (w0 ) =

1+i

1 2

e−(1+2i)f (w0 ) ,

F30 (w0 ) = 0,

e−(1+i)f (w0 ) , 1 + 2i

G20 (w0 ) =

e−(1+i)f (w0 ) , 2w0 G30 (w0 ) = 0, G03 (w0 ) = 0, G21 (w0 ) = −

1

F20 (w0 ) = −

1 − 2i 2w0

2w0

e−(1+2i)f (w0 ) ,

e−(1−2i)f (w0 ) ,

G12 (w0 ) = −

1−i 2w0

e−(1−i)f (w0 ) ,

to derive H00 (w0 ) = 0, H01 (w0 ) =

1 2w0 1

H20 (w0 ) = −

e−f (w0 ) ,

2w0

H30 (w0 ) = 0, H12 (w0 ) =

1

H10 (w0 ) =

H11 (w0 ) = −

e−(1+3i)f (w0 ) ,

2w0

2w0

e−f (w0 ) ,

3+i 2w0

e−(1+i)f (w0 ) ,

H02 (w0 ) = −

2+i

H21 (w0 ) =

3 − 2i

e−(1+2i)f (w0 ) ,

2w0

2−i 2w0

e−(1−i)f (w0 ) ,

e−(1+2i)f (w0 ) ,

H03 (w0 ) =

1−i 2w0

e−(1−2i)f (w0 ) .

Therefore, Z (g11 )(z ,w0 ) =

1 − 2i 2w0

e−(1+i)f (w0 ) r 2 + O(r 3 ).

(43)

By (42), (43) and (20) we then obtain 1 Γ11 (z , w0 ) =

1 − 2i

w0

e−if (w0 ) + O(|z |)

for any z ∈ ∂ D(eif (w0 ) ). Next, note that 1 Γ11 (z , w0 )(ρz ρz w − ρw )(z ,w0 ) = −(2 + i) e−(1+i)f (w0 ) + O(r),

(44)

[ ] ρz w (ρz ρz w − ρw ) − |ρz |2 ρz ww + 2ρz (ρww − |ρz w |2 ) (z ,w

(45)

and 0)

−(1+i)f (w0 )

= 2e

+ O(r),

and substitution from (44)–(46) and ρz (z , w0 ) = re−iϕ − e−if (w0 ) into (21) gives 1 2i re−iϕ − e−if (w0 ) g11 (z , w0 )Γ01 (z , w0 ) = −i e−(1+i)f (w0 ) + O(r).

{

}

Let us multiply by ρz (γ (ϕ )) = reiϕ − ei f (w0 ) and take into account that |ρz (γ (ϕ ))| = 1. Then 1 g11 (z , w0 ) Γ01 (z , w0 ) =

1 2|w0 |2

+ O(r)

(46)

for any z = reiϕ ∈ D(eif (w0 ) ). In order to compute the curvature from (39), we use

{ (

⏐2 }

1 Z (Γ11 ) = (g 11 )2 Z Z (g11 ) g11 − ⏐Z (g11 )⏐

We need the following

)



.

(47)

E. Barletta et al. / Journal of Geometry and Physics 120 (2017) 142–168

153

Lemma 3. For any z ∈ D(eif (w0 ) ) we have that

(

)

Z Z (g11 )

(z ,w0 )

7+i

=

2|w0 |4

r 2 + O(r 3 ).

(48)

Proof. We observe that

(

Kα1 α 2 (w )z α1 z



)

Z Z (g11 ) =

α2

(49)

0≤|α|≤4

where, simply computations give that K00 (w ) = −

1

w

η′ (f (w))H01 (w ) − eif (w)

∂ H00 , ∂w

∂ H00 ∂H − eif (w) 10 , ∂w ∂w 2 ′ i if (w) if (w ) ∂ H01 H01 (w ) − e , K01 (w ) = − η (f (w ))H02 (w ) + e w w ∂w 1 i ∂ H10 ∂H K20 (w ) = − η′ (f (w ))H21 (w ) − e−if (w) H11 (w ) + − eif (w) 20 , w w ∂w ∂w K10 (w ) = −

K11 (w ) = −

1

w

2

w

η′ (f (w))H11 (w ) −

η′ (f (w))H12 (w ) −

i

e−if (w) H01 (w ) +

w

2i

w

e−if (w) H02 (w ) +

i

w

eif (w) H11 (w )

∂ H01 ∂H − eif (w) 11 , ∂w ∂w 3 ′ 2i if (w) if (w ) ∂ H02 H02 (w ) − e , K02 (w ) = − η (f (w ))H03 (w ) + e w w ∂w i ∂ H20 ∂H K30 (w ) = − e−if (w) H21 (w ) + − eif (w) 30 , w ∂w ∂w 2i −if (w) i if (w) ∂ H11 ∂H K21 (w ) = − e H12 (w ) + e H21 (w ) + − eif (w) 21 , w w ∂w ∂w 3i −if (w) 2i if (w) ∂ H02 if (w ) ∂ H12 K12 (w ) = − e H03 (w ) + e H12 (w ) + −e , w w ∂w ∂w 3i ∂ H03 K03 (w ) = eif (w) H03 (w ) − eif (w) . w ∂w +

In order to check (48), notice that by the first part of Lemma 3 K00 (w0 ) = 0, K01 (w0 ) =

1+i 2

e−(2−i)f (w0 ) ,

K11 (w0 ) = −

3i 2

e−2f (w0 ) ,

K10 (w0 ) =

1+i

2 3 + 2i

e−(2+i)f (w0 ) ,

e−2(1+i)f (w0 ) , 2 3 + i −(2−i)f (w0 ) K02 (w0 ) = − e . 2

K20 (w0 ) = −

Substitution into (49) then yields (48). Lemma 3 is proved. □ End of the Proof of Proposition 1. Notice that (43) implies

⏐ ⏐ ⏐Z (g )(z ,w ) ⏐2 = 5 e−3f (w0 ) r 4 + O(r 5 ). 11 0 4

(50)

Next let us substitute into (47) from (42) and (48)–(50) so that to derive 1 Z (Γ11 )(z ,w0 ) =

2+i + O(r). |w0 |2

(51)

Finally (39), (46) and (51) yield R11 (z , w0 ) = −

2

|w0 |2

+ O(r),

(52)

which is (40). Hence (by (42)) R(z , w0 ) = −

4 r2

+ O(r −1 ),

which is (41). Proposition 1 is therefore proved. □

(53)

154

E. Barletta et al. / Journal of Geometry and Physics 120 (2017) 142–168

A complex valued differential p-form ω on a CR manifold (N , T1,0 (N)) has type (p, 0) if T0,1 (N) ⌋ ω = 0. Let Λp,0 (N) → N be the bundle of all (p, 0)-forms. The canonical bundle K (N) → N is the complex line bundle K (N) = Λn+1,0 (N), the top degree (p, 0)-forms, where n is the CR dimension of N. Next R+ acts on K (N) so that the quotient space C (N) = [K (N) \ {zero section}] /R+ is the total space of a S 1 -bundle over N, called the canonical circle bundle, with projection ΠN : C (N) → N. Let W ⊂ C2 be a worm domain and let C (∂ W ) be the canonical S 1 -bundle over ∂ W , with projection Π = Π∂ W . C (∂ W ) carries a natural foliation V by circles, called the vertical foliation, tangent to the distribution Ker(dΠ ). We also recall that a subset of a foliated manifold is called saturated if it is a union of leaves. Lemma 4. It holds that C (M) = Π −1 (M). In particular C (M) is an open saturated subset of (C (∂ W ), V ). Recall that we set M = ∂ W \ A. By construction, the metric Fθ is (cf. Definition 2.15 in [19], p. 128)

˜ θ + 2 (π ∗ θ ) ⊙ σ , Fθ = π ∗ G σ =

1{

d s + π ∗ σ0 ,

3

}

(54)

σ0 = i ω1 1 −

i 2

g 11 dg11 −

R 8

θ,

(55)

where s is a local fiber coordinate on C (M). Also ω1 1 , g11 and R are respectively the connection 1-forms of the Tanaka–Webster connection of the pseudohermitian manifold (M , θ ), the Levi invariant, and the pseudohermitian scalar curvature. Let M be a 4-dimensional manifold. A Lorentzian metric on M is a nondegenerate symmetric smooth (0, 2)-tensor field F on M of signature (− + + +). A tangent vector v ∈ Tp (M) is spacelike, resp. timelike, resp. null, if Fp (v, v ) > 0, resp. if Fp (v, v ) < 0, resp. v ̸ = 0 and Fp (v, v ) = 0. A vector field X , defined on some open subset U ⊂ M is called timelike if Xp is a timelike tangent vector for each p ∈ U . A globally defined timelike vector field X on M is referred to as a time orientation of the Lorentzian manifold (M, F ) and the synthetic object (M, F , X ) is a space–time. A tangent vector v ∈ Tp (M) is future, resp. past, directed if Fp (v, Xp ) < 0, resp. if Fp[(v,( Xp ) >)0. ] The pointwise restriction of θ = ι∗ 2i ∂ − ∂ ρ to M is a positively oriented contact form on M. C.R. Graham recognized (cf. [20]) σ as a connection 1-form in the principal bundle S 1 → C (M) → M. Let X ↑ ∈ X(C (M)) denote the horizontal lift of X ∈ X(M) with respect to σ i.e.

σ (X ↑ ) = 0, (dp π )Xp↑ = Xπ (p) , p ∈ C (M). If S ∈ X(C (M)) is the tangent to the S 1 action and T ∈ X(M) the Reeb vector field of (M , θ ) then Xθ ≡ T ↑ − S is a time orientation of (C (M), Fθ ), hence (C (M), Fθ , Xθ ) is a space–time. Let K be the scalar curvature of Fθ . As S 1 ⊂ Isom(C (M), Fθ ) it follows that K is S 1 -invariant. Hence, there is a unique function π∗ K ∈ C ∞ (C (M), R) whose vertical lift is K . By a result of J.M. Lee, [13] (cf. also [19], p. 142)

π∗ K = 3R/2.

(56)

Finally for any lift Γ of γw0 lim

ϕ→f (w0 )+π /2

K (Γ (ϕ )) =

3 2

lim

ϕ→f (w0 )+π /2

R(γw0 (ϕ )) = −∞

as a consequence of Proposition 1. Theorem 1 is proved. 4. Existence of timelike lifts To start with one produces a local coordinate neighborhood (U , xj ) on M such that γw0 (ϕ ) ∈ U for any 0 ≤ ϕ−f (w0 ) < π/2 i.e. Lemma 5. The set U =

{(

r ei ϕ (r ,w) , w : r > 0, w ∈ C∗ , r 2 − 2r + η(f (w )) < 0 ,

)

ϕ (r , w) ≡ f (w) + arccos

}

[

r 2

+

η(f (w)) 2r

]

,

(57) (58)

is an open subset of M and the map

χ : U → (0, +∞) × C∗ ,

χ (z , w) = (|z | , w ), (z , w) ∈ U ,

(59)

is a homeomorphism on χ (U) with the inverse

( ) χ −1 (r , w) = r ei ϕ (r ,w) , w ,

(r , w ) ∈ χ (U).

The curve γ = γw0 lies in U i.e. γ (ϕ ) ∈ U for any 0 ≤ ϕ − f (w0 ) <

π 2

.

E. Barletta et al. / Journal of Geometry and Physics 120 (2017) 142–168

155

Proof. Let us set z = reiϕ . Then ρ (z , w ) = 0 yields cos [ϕ − f (w )] =

r

+

2

η(f (w)) 2r

where one may solve for ϕ provided that r 2 − 2r + η(f (w )) ≤ 0 so that ϕ = ϕ (r , w ) where ϕ (r , w ) is given by (58). The construction of the fiber coordinate s in (55) may be detailed as follows. For an arbitrary fixed s0 ∈ R

λ , |λ|

|s − s0 | < π, [ ( ) ] p = λ θ ∧ θ 1 x , x = (z , w ) ∈ U , ( λ ) Φ : C (M) → M × S 1 , Φ (p) = x , , |λ| U ≡ Φ −1 (M × V (s0 )) ⊂ M, V (s0 ) = {eis : |s − s0 | < π}. s : U → (s0 − π , s0 + π ) ,

s(p) = s,

eis =

The local coordinates corresponding to the local chart χ in Lemma 5 are denoted by

) ( χ = (r , w) = x˜ 1 , x˜ 2 + i x˜ 3 . Let xj = x˜ j ◦ π : U → R with U = Φ −1((U × V (0)) so)that (U , xj , s) is an induced local coordinate system on C (M). We often set x0 = s so that {xα : 0 ≤ α ≤ 3} ≡ s, x1 , x2 , x3 . Lemma 6. Let us set θ 1 = β dz + α dw where

α=

1

ρz − H ρw

,

β = Hα, H =

ρw − ρz ρz w . ρw ρz w − ρz ρww

Then θ 1 (Z ) = 1, θ 1 (Z ) = 0 and θ 1 (T ) = 0 so that {θ 1 , θ 1 , θ } is an admissible frame of T ∗ (M) ⊗ C. Here θ 1 = θ 1 . The proof is a straightforward calculation. Any lift Γ of γ = γw0 may be written as

[ ( ) ] Γ (ϕ ) = λ(ϕ ) θ ∧ θ 1 γ (ϕ ) for some C ∞ function λ : I(w0 ) → C. Let us set

Γ α (ϕ ) = xα (Γ (ϕ )),

0 ≤ α ≤ 3,

so that eiΓ

0 (ϕ )

=

λ(ϕ ) , Γ j (ϕ ) = γ j (ϕ ), |λ(ϕ )|

1 ≤ j ≤ 3,

where γ j = xj ◦ γ . Lemma 7. The tangent vector to γ = γw0 may be represented as

γ˙ (ϕ ) = 2 Tγ (ϕ ) −

1

1

2 cos2 [ϕ − f (w0 )]

{

w0 ei f (w0 ) Z + w0 e−i f (w0 ) Z

}

γ (ϕ )

.

(60)

Consequently the tangent vector to any lift Γ : I(w0 ) → C (M) of γ is given by

[

Γ˙ (ϕ ) = γ˙ (ϕ )↑ +

dΓ 0 dϕ

](

(ϕ ) + σ0, γ (ϕ ) (γ˙ (ϕ ))

∂ ∂s

) (61) Γ (ϕ )

where horizontal lifting is meant with respect to σ . The proof of Lemma 7 follows by a rather involved evaluation of the first and second order derivatives of ρ along the circle

{ } γ : z = eif (w0 ) 1 + e2i[ϕ−f (w0 )] ,

w = w0 ,

(62)

that is,

ρz (γ (ϕ )) = e−i[2ϕ−f (w0 )] , } i { 2i[ϕ−f (w0 )] ρw (γ (ϕ )) = e − e−2i[ϕ−f (w0 )] , w0 ρz w (γ (ϕ )) = ρww (γ (ϕ )) =

i

w0

e−i f (w0 ) ,

1

|w0 |

)2

ei[ϕ−f (w0 )] + e−i[ϕ−f (w0 )] .

( 2

(63)

156

E. Barletta et al. / Journal of Geometry and Physics 120 (2017) 142–168

Also if G = 1/H then β = 1/(G ρz − ρw ) hence, using (63), G(γ (ϕ )) = −

β (γ (ϕ )) =

i

eiϕ e2i[ϕ−f (w0 )] ei[ϕ−f (w0 )] + e−i[ϕ−f (w0 )]

{

w0

}

{ } × e2i[ϕ−f (w0 )] + e−2i[ϕ−f (w0 )] , e−2i[ϕ−f (w0 )]

iw0

(64)

.

4 cos2 [ϕ − f (w0 )]

The tangent vector along (62) is given by

} { γ˙ (ϕ ) = 2i ei[2ϕ−f (w0 )] (∂z )γ (ϕ ) − e−i[2ϕ−f (w0 )] (∂z )γ (ϕ ) .

(65)

Let us apply

θ=

i 2

(ρz dz − ρz dz + ρw dw − ρw dw)

(66)

to (65) and take into account the first equation in (63). We obtain

θγ (ϕ ) (γ˙ (ϕ )) = 2.

(67)

Similarly, one applies θ 1 = β dz + α dw to (65) so that to get

θγ1(ϕ ) (γ˙ (ϕ )) = 2i β (γ (ϕ )) ei[ϕ−f (w0 )] and then, using (64),

θγ1(ϕ ) (γ˙ (ϕ )) = −

ei f (w0 )

w0

2 cos2 [ϕ − f (w0 )]

.

(68)

Finally (67) and (68) yield (60). In order to prove (61) one starts from

Γ˙ (ϕ ) =

dΓ 0 dϕ

(ϕ )

(

∂ ∂s

) + Γ (ϕ )

dγ j dϕ

(ϕ )

(

∂ ∂ xj

) Γ (ϕ )

.

(69)

On the other hand, by taking into account the direct sum decomposition TΓ (ϕ ) (C (M)) = Ker(σ )Γ (ϕ ) ⊕ Ker dΓ (ϕ ) π ,

(

)

one may represent Γ˙ (ϕ ) as ↑ Γ˙ (ϕ ) = VΓ (ϕ ) + µ(ϕ ) SΓ (ϕ ) ,

(70)

for some tangent vector field V ∈ X(M) and some smooth function µ : I(w0 ) → R. Substitution from (69) into (70) followed by applying σΓ (ϕ ) yields Vγ (ϕ ) = γ˙ (ϕ ),

µ(ϕ ) =

dΓ 0 dϕ

(ϕ ) + σ0, γ (ϕ ) (γ˙ (ϕ ))

and Lemma 7 is proved. □ Theorem 3 is an immediate corollary to (60) in Lemma 7. Indeed gθ , γ (ϕ ) (γ˙ (ϕ ), γ˙ (ϕ )) = 4 +

1 cos2

[ϕ − f (w0 )]

hence E(γ ) = +∞. ˜ θ = Gθ Next, we need to compute the length of Γ˙ (ϕ ) with respect to the Fefferman metric Fθ . To this end we recall that G ˜ on H(M) ⊗ H(M) and G(T , V ) = 0 for any V ∈ X(M). Also, Gθ (Z , Z ) = 0,

Gθ (Z , Z ) = g11 ,

hence, by Lemma 7,

( ) π ∗ G˜ θ

Γ (ϕ )

( ) |w0 |2 Γ˙ (ϕ ) , Γ˙ (ϕ ) = 2

g11 (γ (ϕ )) cos4 [ϕ − f (w0 )]

.

(71)

Also (as previously shown)

( ∗ ) π θ Γ (ϕ ) Γ˙ (ϕ ) = 2,

(72)

E. Barletta et al. / Journal of Geometry and Physics 120 (2017) 142–168

σΓ (ϕ ) Γ˙ (ϕ ) =

1

[

dΓ 0 dϕ

3

157

]

(ϕ ) + σ0,γ (ϕ ) (γ˙ (ϕ )) .

(73)

The formulae (71)–(73), together with (54) yield Fθ ,Γ (ϕ ) Γ˙ (ϕ ) Γ˙ (ϕ ) =

(

)

g11 (γ (ϕ ))

|w0 |2

[[ϕ −0f (w0 )] ] dΓ + (ϕ ) + σ0,γ (ϕ ) (γ˙ (ϕ )) . 3 dϕ cos4

2

(74)

4

As a byproduct of the proof of Proposition 1 we obtain g11 (z , w0 ) =

1 2

|z | e−f (w0 ) ,

hence g11 (γ (ϕ )) =

2

|w0 |2

cos2 [ϕ − f (w0 )] .

(75)

By (74) and (75), it follows that Γ is timelike if and only if 4

[

dΓ 0 dϕ

3

]

(ϕ ) + σ0,γ (ϕ ) (γ˙ (ϕ ))

cos2 [ϕ − f (w0 )] < −1.

(76)

Finally, by integrating in (76), if g : I(w0 ) → R is an arbitrary strictly decreasing smooth function,

Γ (ϕ ) = g(ϕ ) − 0

3 4

tan [ϕ − f (w0 )] −

ϕ



f (w0 )

σ0,γ (t) (γ˙ (t)) dt

(77)

then Γ is timelike. For any lift Γ of γ given by (77), by (76), one has Fθ , Γ (ϕ )

(

] [ ) 1 dΓ 0 ˙ Γ (ϕ ) , Xθ , Γ (ϕ ) = −1 + (ϕ ) + σ0,γ (ϕ ) (γ˙ (ϕ )) < 0 3 dϕ

hence Γ is future directed. Theorem 2 is proved. 5. Levi flat points and space–time singularities ΠL

Let M = C (M) and let GL(4, R) → L(M) −→ M be the principal bundle of all linear frames tangent to M. An element u ∈ L(M) with ΠL (u) = p is an R-linear isomorphism u : R4 → Tp (M). Let ψ ∈ C ∞ (T ∗ (L(M)) ⊗ gl(4, R)) be the Levi-Civita connection 1-form of the Fefferman metric Fθ i.e.

( ) ψ A∗ = A,

R∗a ψ = ad(a−1 ) ψ,

for any left invariant vector field A ∈ gl(4, R) and any a ∈ GL(4, R). Let Θ ∈ C ∞ T ∗ (L(M)) ⊗ R4 be the canonical 1-form i.e.

(

Θu = u−1 ◦ (du ΠL ) ,

)

u ∈ L(M).

The Schmidt metric (cf. [15]) is the Riemannian metric G on L(M) given by G(V , W ) = ψ (V ) · ψ (W ) + Θ (V ) · Θ (W ) for any V , W ∈ T (L(M)) and the ‘‘dot-products’’ are the Euclidean inner products on R16 ≈ gl(4, R) and R4 . As M is oriented, L(M) has two connected components L± (M). Let GL+ (4) be the Lie group consisting of all a ∈ GL(4, R) such that det(a) > 0. Let dG : L+ (M) × L+ (M) → [0, +∞) be the distance function associated to the Riemannian metric G i.e. dG (u, v ) is the infimum of lengths of all piecewise C 1 curves joining u, v ∈ L+ (M). Let L+ (M) be the Cauchy completion of L+ (M) with respect to the distance function dG . For each a ∈ GL+ (4) the right translation Ra : L+ (M) → L+ (M) is uniformly continuous with respect to dG so that the action of GL+ (4) on L+ (M) extends to L+ (M) as a topological group action. Let

M = L+ (M)/GL+ (4) be the orbit space. There is a natural injection

M ≈ L+ (M)/GL+ (4) ↪→ L+ (M)/GL+ (4).

˙ = M \ M. A frame u ∈ L+ (M) with ΠL (u) = p is Lorentzian if The b-boundary of the space–time M is defined as M Fθ , p u(eα ) , u(eβ ) = ϵα δαβ ,

(

ϵ0 = −1 = −ϵj ,

)

1 ≤ j ≤ 3.

0 ≤ α , β ≤ 3,

158

E. Barletta et al. / Journal of Geometry and Physics 120 (2017) 142–168

Let O(1, 3) be the Lorentz group and O(M) → M the principal O(1, 3)-bundle of Lorentizan frames tangent to M. Let Σ : M → L(M) be the section given by (11). Note that Σ (M) ⊂ O(M) i.e. Σ (p) is a Lorentzian ∑n−1 i i frame for any p ∈ M. We adopt the notation Rn1 for Rn together with the Minkowski form η1,n−1 = −x0 y0 + i=1 x y of index ν = 1 [so that O(1, n − 1) consists of all a ∈ GL(n, R) preserving η1,n−1 ]. Each a ∈ O(1, 3) may be represented as

( a=

aT c

b aS

)

where aT : R11 → R11 (the timelike part of a) is the restriction of a to R11 composed with the orthogonal projection on R11 while aS : R3 → R3 (the spacelike part of a) is the restriction of a to R3 followed by orthogonal projection on R3 . We set as customary O++ (1, 3) = {a ∈ O(1, 3) : aT (1) > 0, det(aS ) > 0} , which is a connected component of O(1, 3). Next let us fix a connected component O+ (M) such that O+ (M) is a submanifold of L+ (M) and Σ (M) ⊂ O+ (M). Schmidt’s metric G induces a Riemannian metric on O+ (M) [the first fundamental form of ι : O+ (M) ↪→ L+ (M)] and hence a distance function dι∗ G . By a result in [15] the b-boundary of Fefferman’s space–time M may be recovered as

[

] [

˙ = O+ (M)/O++ (1, 3) \ O+ (M)/O++ (1, 3) M

]

where O+ (M) is the Cauchy completion of O+ (M) with respect to dι∗ G . Let Γ : I(w0 ) → M be an arbitrary lift of the circle γ = γw0 i.e. π ◦ Γ = γ . Let Γ H : I(w0 ) → O+ (M) be a horizontal lift of Γ with respect to the connection 1-form ω ∈ C ∞ (T ∗ (O+ (M)) ⊗ o(1, 3)) [the Levi-Civita connection 1-form of Fθ ] i.e. ΠO+ ◦ Γ H = Γ and Γ˙ H (ϕ ) is a horizontal vector. ( ) Here ΠO+ : O+ (M) → M is the projection. As ωΓ H (ϕ ) Γ˙ H (ϕ ) = 0 the length of Γ H is f (w0 )+π /2



f (w0 )

(

f (w0 )+π /2

∫ =

f (w0 ) f (w0 )+π /2



)1/2

GΓ H (ϕ ) Γ˙ H (ϕ ) , Γ˙ H (ϕ )

= f (w0 )



 ( ) ΘΓ H (ϕ ) Γ˙ H (ϕ )  dϕ  H −1  Γ (ϕ ) Γ˙ (ϕ ) dϕ,

where ∥ξ ∥ is the Euclidean norm of ξ ∈ R4 . Together with a result in [15] this yields

˙ if and only Theorem 7. The limit limϕ→log |w0 |2 +π /2 Γ (ϕ ) exists in the M topology and determines a point on the b-boundary M if the improper integral π /2



 H  Γ (t + log |w0 |2 )−1 Γ˙ (t + log |w0 |2 ) dt

0

is convergent. However one may exhibit another class of lifts of Γ to O+ (M) reaching their end point in infinite proper time. Precisely let C : I(w0 ) → O+ (M) be given by C = Σ ◦ Γ . Let us endow M, M and L+ (M) respectively with )the local coordinate ( systems (U , x˜ i ), (U , xµ ) [induced by (U , x˜ i ) i.e. xi = x˜ i ◦ π and x0 = s, cf. Section 4] and ΠL−1 (U , X µ , Xνµ [induced by (U , xµ ) i.e. X µ = xµ ◦ ΠL and Xνµ are the fiber coordinates]. Let C µ (ϕ ) = X µ (C (ϕ )) = Γ µ (ϕ ),

Cνµ (ϕ ) = Xνµ (C (ϕ )),

be the local components of C . If V is a tangent vector field on M then V H denotes its horizontal lift with respect to ω i.e. VuH ∈ Ker(ωu ),

(du ΠO+ )VuH = VΠO+ (u) ,

u ∈ O+ (M).

Also if v ∈ Tp (M) and u ∈ O+ (M) is a Lorentzian frame with origin at ΠO+ (u) = p then the horizontal lift of v is v H = βu (v ) where βu is the R-linear isomorphism

[ ]−1 βu ≡ du ΠO+ : Ker(ωu ) → Tp (M) . The horizontal lift of ∂/∂ xα is given by (cf. e.g. [21], vol. I)

∂ ∂ xα { }

(

)H =

∂ ∂Xα

} ∂ µ Xβ αβ ν ∂ Xνµ

{ −

(78)

µ

where αβ are the Christoffel symbols of gµν = Fθ (∂µ , ∂ν ) with ∂µ ≡ ∂/∂ xµ . Then, (78) implies

[ µ { } ]( ) α ∂ ˙C (ϕ ) = Γ˙ (ϕ )H + dCν (ϕ ) + dΓ (ϕ ) µ (Γ (ϕ ))Cνβ (ϕ ) . µ αβ dϕ dϕ ∂ Xν C (ϕ )

(79)

E. Barletta et al. / Journal of Geometry and Physics 120 (2017) 142–168

159

Let us set

) ( ξ (ϕ ) = ΘC (ϕ ) C˙ (ϕ ) ∈ R4 . If ξ (ϕ ) = ξ α (ϕ )eα then

[ ] Γ˙ (ϕ ) = ξ α (ϕ ) Eα (Γ (ϕ )) = −ξ 0 (ϕ ) + ξ 3 (ϕ ) SΓ (ϕ ) +

{[

(80)

] }↑ ξ 0 (ϕ ) + ξ 3 (ϕ ) T + ξ 1 (ϕ ) E1 + ξ 2 (ϕ ) E2 Γ (ϕ ) .

Here S is given by (10) in Section 1. Also note that

ξ 1 (ϕ ) E1,γ (ϕ ) + ξ 2 (ϕ ) E2,γ (ϕ ) = √

1

[

2g11

ζ (ϕ ) Zγ (ϕ ) + ζ (ϕ ) Z γ (ϕ )

]

(81)

where ζ = ξ 1 + i ξ 2 . Then, by (61), (80) and (81), we have 2 3

[

dΓ 0 dϕ

] (ϕ ) + σ0,γ (ϕ ) (γ˙ (ϕ )) = −ξ 0 (ϕ ) + ξ 3 (ϕ ),

and

[ ] γ˙ (ϕ ) = ξ 0 (ϕ ) + ξ 3 (ϕ ) Tγ (ϕ ) + √

1

[

2g11 (γ (ϕ ))

] ζ (ϕ ) Zγ (ϕ ) + ζ (ϕ ) Z γ (ϕ ) ,

yielding, using (60) and (75),

ζ (ϕ ) = −

w0 eif (w0 ) , |w0 | cos [ϕ − f (w0 )]

(82)

and

ξ 3,0 (ϕ ) = 1 ±

1

[

dΓ 0 dϕ

3

(ϕ ) + σ0,γ (ϕ ) (γ˙ (ϕ ))

] (83)

by (82)–(83) we may conclude that

  ΘC (ϕ ) C˙ (ϕ )2 = 2 +

2 dΓ 0

[

1 cos2 [ϕ − f (w0 )]

+

9



]2 (ϕ ) + σ0,γ (ϕ ) (γ˙ (ϕ )) .

Finally,



f (w0 )+π /2 f (w0 )

GC (ϕ )

)1/2 C˙ (ϕ ), C˙ (ϕ ) dϕ ≥

(



f (w0 )+π /2 f (w0 )



f (w0 )+π /2

≥ f (w0 )

  ΘC (ϕ ) C˙ (ϕ ) dϕ dϕ cos [ϕ − f (w0 )]

= ∞. Theorem 4 is then proved. 6. Dimension reduction and b-boundary points Section 6 is devoted to a dimension reduction argument, leading to the determination of subsets of the adapted bundle boundary ∂adt M, that we will introduce shortly. The approach is inspired by the work of R.A. Johnson [17] and B. Bossard [18]. Most of the theoretical considerations in Section 6, such as induced and normal connections, bundle boundary constructions, etc., apply to any Lorentzian surface Σ in (M, Fθ ) such that the immersion Σ ↪→ M has a flat normal connection. Natural examples to keep in mind are saturated subsets of M got as vertical lifts of circles γw0 i.e. total spaces of principal circle subbundles S 1 → Σ → Γw0 . Precisely for each |f (w0 )| ≤ µ we set

{ } Γw0 = γw0 (ϕ ) : ϕ ∈ I(w0 ) \ {f (w0 )} . ( ) Then Σ = π −1 Γw0 ≈ (0 , π2 ) × S 1 is a real hypersurface in M.(It will be useful to represent Σ as the translation by the ) S 1 -action of the σ -horizontal lift of γ = γw0 . To this end, let x0 = 2eif (w0 ) , w0 ∈ M so that γ issues at x0 . Let p0 ∈ π −1 (x0 ) and let

γ ↑ = γ ↑ ( · ; p0 ) : I(w0 ) → M be the σ -horizontal lift of γ issuing at p0 .

160

E. Barletta et al. / Journal of Geometry and Physics 120 (2017) 142–168

Lemma 8. (i) Set

} { Σ (p0 ) = γ ↑ (ϕ; p0 ) · ζ : ϕ ∈ I(w0 ) \ {f (w0 )}, ζ ∈ S 1 . Then Σ = Σ (p0 ). (ii) Σ is a Lorentzian surface with the induced metric ι∗ Fθ , where ι : Σ ↪→ M. In particular if T (Σ )⊥ → Σ is the normal bundle of ι then Fθ ,p is positive definite on T (Σ )⊥ p . (iii) For any point p = γ ↑ (ϕ ) · ζ ∈ Σ

{ |w0 | X↑ + cos [ϕ − f (w0 )] { |w0 | N2 (p) = Y↑ + cos [ϕ − f (w0 )]

N1 (p) =

1

[

1

e−if (w0 ) +

2 w0 [ 1 i 2

e−if (w0 ) −

w0

1

w0 1

w0

] }

eif (w0 ) S

] }p

eif (w0 ) S

, (84)

,

p

is an orthonormal basi in the normal space at p, that is, Fθ (Ni , Nj ) = δij for 1 ≤ i, j ≤ 2. Proof. (i) The proof is straightforward. Note also that given q0 ∈ π −1 (x0 ) there is ζ ∈ S 1 such that q0 = p0 · ζ and hence

γ ↑ (ϕ; q0 ) = γ ↑ (ϕ; p0 ) · ζ . (ii) For every 0 < ϕ − f (w0 ) < π2 the tangent vectors γ˙ ↑ (ϕ ) and Sγ ↑ (ϕ ) are linearly independent. Indeed γ˙ ↑ (ϕ ) ∈ Ker(σ )γ ↑ (ϕ ) and Sγ ↑ (ϕ ) ∈ Ker(dγ ↑ (ϕ ) π ) and Ker(σ ) ∩ Ker(dπ ) = (0). Let p = γ ↑ (ϕ ) · a ∈ Σ and V ∈ Tp (Σ ). Then V = λ γ˙ ↑ (ϕ ; p0 · a) + µ Sp

(85)

for some λ, µ ∈ R. Assume that Fθ ,p (V , W ) = 0 for any W ∈ Tp (Σ ). Then, 0 = Fθ ,p V , γ˙ ↑ (ϕ ; p0 · a)

(

)

= λ G˜ θ ,γ (ϕ ) (γ˙ (ϕ ) , γ˙ (ϕ )) + µ θγ (ϕ ) (γ˙ (ϕ )) σ (S)p λ + µ, = cos2 [ϕ − f (w0 )] and 0 = Fθ ,p (V , Sp ) = (λ/2) θγ (ϕ ) (γ˙ (ϕ )) = λ, yield λ = µ = 0. It follows that (ι∗ Fθ )p is nondegenerate for any p ∈ Σ . In order to see that the scalar product (ι∗ Fθ )p has index 1 it suffices to produce a timelike vector in Tp (Σ ). Indeed if V is given by (85) then

λ2 + 2λµ [ϕ − f (w0 )] hence Fθ ,p (V , V ) < 0 for suitably chosen λ and µ. Fθ ,p (V , V ) =

cos2

(iii) With the notations in Section 1 one may look for N ∈ C ∞ (T (Σ )⊥ ) as N = λa Ea↑ + µ+ T ↑ + S + µ− T ↑ − S

(

(

)

)

for some λa , µ± ∈ C ∞ (Σ , R). The tangent space Tp (Σ ) is the span of γ˙ ↑ (ϕ; p0 · ζ ), Sp . Then

{

Fθ ,p (N , S)p = 0,

µ+ + µ− = 0, √ µ+ − µ− =

Fθ ,p (Np , γ˙ ↑ (ϕ; p0 · ζ )) = 0,

g11

1

2 2 cos2 [ϕ − f (w0 )]

{

} w0 β eif (w0 ) + w0 β e−if (w0 ) ,

with β = λ1 − i λ2 . Therefore,

{ } 1 −i(f (w0 )) β |w0 | ↑ N = Z + e S 2 cos[ϕ − f (w0 )] w0 { } 1 i(f (w0 )) β |w0 | ↑ + Z + e S 2 cos[ϕ − f (w0 )] w0 and hence N = 2λa Na where Na are given by (84). □ Let t ∈ R and let

( Bt =

cosh t sinh t

sinh t cosh t

)

}

E. Barletta et al. / Journal of Geometry and Physics 120 (2017) 142–168

161

be the boost of R21 through (oriented) Lorentz angle t. Then O++ (1, 1) = {Bt : t ∈ R} is the component of the identity in the Lorentz group O(1, 1). Let ι : Σ ↪→ M be a Lorentzian surface and let O(1, 1) →

↪→

O(Σ ) πO ↓ Σ

← GL(2, R)

L(Σ )

↓ πL Σ

(86)

be the subbundle of Lorentzian frames tangent to (Σ , ι∗ Fθ ). Let O+ (Σ ) be a connected component of O(Σ ). To consider the πO+

Schmidt distance function we shall work with the principal bundle O++ (1, 1) → O+ (Σ ) → Σ with πO+ = πO |O+ (Σ ) rather than (86). Let

( ) ψ T ∈ C ∞ T ∗ (L(Σ )) ⊗ gl(2, R) ,

( ) ΘΣ ∈ C ∞ T ∗ (L(Σ )) ⊗ R2 ,

be respectively the Levi-Civita connection 1-form of the Lorentzian [by (ii) in Lemma 8] ι∗ Fθ and the canonical 1-form

ΘΣ ,u = u−1 ◦ (du πL ),

u ∈ L(Σ ).

Clearly ψ is reducible to a connection 1-form T

( ) ψ T ∈ C ∞ T ∗ (O+ (Σ )) ⊗ o(1, 1) . Let GΣ = ψ T · ψ T + ΘΣ · ΘΣ be the Schmidt metric on L(Σ ). The same symbol GΣ denotes the induced Riemannian metric on O+ (Σ ). Let dΣ : O+ (Σ ) × O+ (Σ ) → [0, +∞) be the distance function associated to GΣ and let O+ (Σ ) be the Cauchy completion of O+ (Σ ) with respect to dΣ . We set as usual

Σ = O+ (Σ )/O++ (1, 1),

˙ = Σ \ Σ. Σ

We identify O(1, 1) and O(2) with subgroups of O(1, 3) i.e. O(1, 1) ≈

O(1, 1) 0

(

0 I2

)

(

,

O(p) ≈

)

I1,1 0

0 , O(2)

where I2 is the identity matrix of order 2 and I1,1 = diag (−1, 1). Let us set O(M)|Σ = {v ∈ O(M) : ΠO (v ) ∈ Σ } where ΠO : O(M) → M is the projection. Let O(1, 1) × O(2) → O(M, Σ ) → Σ be the principal bundle of adapted frames i.e. O(M, Σ ) consists of all v ∈ O(M)|Σ such that v : R4 → Tp (M) maps R2 × {0} onto Tp (Σ ). We shall need the bundle map hT : O(M) → O(Σ ),

hT (v ) = v ⏐R2 ×{0} .



Let O(2) → O(T (Σ )⊥ ) ≈ O(M, Σ )/O(1, 1) → Σ be the principal bundle of all normal frames v : R2 → T (Σ )⊥ p with p ∈ Σ and let us consider the bundle map h⊥ : O(M, Σ ) → O(T (Σ )⊥ ),

h⊥ (v ) = v ⏐{0}×R2 .



We collect the various bundles and bundle maps introduced so far in the following commutative diagrams

O(Σ ) =

O(1, 1)

O(1, 1) × O(2)

↓ O(M, Σ )



O(2)

hT

O(1, 3)



↓ i

↓Π Σ

O(M)|Σ

↓ j

↓Π =

Σ

= O(1, 3)

−→

O(M)

↓Π ι

↪→

O(M, Σ ) O(1, 1)

↓ Π⊥

Σ

=

−→

−→

↓Π

O(1, 1) × O(2) O(M, Σ )



h⊥

O(M, Σ )

←−

↓ ΠT Σ

O(2)

M

Σ

= O(T (Σ )⊥ )

162

E. Barletta et al. / Journal of Geometry and Physics 120 (2017) 142–168

where i and j are injections. If ΘΣ and Θ are the canonical 1-forms on O(Σ ) and O(M) respectively then (hT )∗ ΘΣ coincides with the restriction of Θ to O(M, Σ ) and in particular the restriction of the R4 -valued form Θ to O(M, Σ ) is R2 -valued [cf. Proposition 1.1 in [21], Vol. 2, p. 3]. Let g be the orthogonal complement of o(1, 1) ⊕ o(2) in o(1, 3) with respect to the Killing–Cartan form. Then the o(1, 1) ⊕ o(2)-component

( ) ω = i∗ j∗ ψ o(1,1)⊕o(2) of i∗ j∗ ψ is a connection 1-form on O(M, Σ ) [cf. Proposition 1.2 in [21], Vol. 2, p. 3]. We also recall [cf. Propositions 1.3 and 1.4 in [21], Vol. 2, p. 4] that: (i) hT : O(M, Σ ) → O(Σ ) maps the connection-distribution defined by ω on O(M, Σ ) onto the Levi-Civita connectiondistribution of (Σ , ι∗ Fθ ); (ii) the Levi-Civita connection 1-form ψ T is determined by (hT )∗ ψ T = ωo(1,1) ;

(87)

(iii) there is a unique connection 1-form ψ



on O(T (Σ ) ) such that ⊥

(h⊥ )∗ ψ ⊥ = ωo(2) ;

(88)

(iv) the map (h , h ) : O(M, Σ ) → O(Σ ) × O(T (Σ ) ) induces a principal bundle isomorphism O(M, Σ ) ≈ O(Σ ) + O(T (Σ )⊥ ) and T





ω = (hT )∗ ψ T + (h⊥ )∗ ψ ⊥ .

(89)

The connection 1-form ω may be used to produce the Schmidt like metric Gadt = ω · ω + Θ · Θ and therefore a bundle completion and bundle boundary

M

adt

= O+ (M, Σ )

adt

/O++ (1, 3),

∂adt M = M

adt

\ M,

adt

where O+ (M, Σ ) is the Cauchy completion of O+ (M, Σ ) with respect to the distance function dadt associated to the Riemannian metric Gadt (rather than i∗ dG i.e. the distance function induced on O+ (M, Σ ) by dG ). We refer to ∂adt M as the adapted bundle boundary of (M, Fθ ). H. Friedrich examined (cf. [22]) b-boundary constructions relying on arbitrary G-structures [with G ⊂ GL(4, R), a Lie subgroup]. By a result in [22] the same b-boundary is produced provided that ψ is reducible to a connection 1-form on the given G-structure. Since i∗ j∗ ψ is not a connection 1-form on O(M, Σ ) it follows ˙ and ∂adt M are in general logically unrelated objects. Our further constructions will require a pair of globally defined that M cross sections sT : Σ → O(Σ ),

s⊥ : Σ → O(T (Σ )⊥ ),

whose existence is postulated. For instance if Σ = π −1 (Γw0 ) then sT (p) : R2 → Tp (Σ ),

s⊥ (p) : R2 → T (Σ )⊥ p ,

sT (p)(e0 ) = (dγ ↑ (ϕ ) Rζ )γ˙ ↑ (ϕ ), s⊥ (p)ej+1 = Nj (p),

sT (p)e1 = Sp ,

j ∈ {1, 2},

for any p = γ (ϕ ) · ζ ∈ Σ . Our main tool in the present section is the map ↑

j : O(Σ ) → O(M, Σ ), j(u)eα = u(eα ),

j(u) : R4 → Tp (M),

j(u)ei+1 = s⊥ (p)ei+1 , α ∈ {0, 1}, i ∈ {1, 2},

for any u ∈ O(Σ )p and any p ∈ Σ . Lemma 9. Let Σ be a Lorentzian surface in (M, Fθ ) endowed with two globally defined sections sT : Σ → O(Σ ),

s⊥ : Σ → O(T (Σ )⊥ ).

If j is an isometry of (O(Σ ), GΣ ) into (O(M, Σ ), Gadt ) then: (i) j is uniformly continuous as a map of metric spaces (O+ (Σ ), dΣ ) and (O+ (M, Σ ), dadt ). adt (ii) The induced map of completions j : O+ (Σ ) → O+ (M, Σ ) is Lipschitz with Lipschitz constant L = 1 [with respect to the induced distance functions dΣ and dadt ] and equivariant i.e. it commutes with the topological group actions of O++ (1, 1) and O++ (1, 3).

E. Barletta et al. / Journal of Geometry and Physics 120 (2017) 142–168

163

Proof. (i) For any pair of Lorentzian frames u′ , u′′ ∈ O+ (Σ ) let ΩΣ (u′ , u′′ ) [respectively ΩM (j(u′ ), j(u′′ ))] be the set of all piecewise C 1 curves in O+ (Σ ) [respectively in O+ (M, Σ )] joining u′ and u′′ [respectively j(u′ ) and j(u′′ )]. Then dadt (j(u′ ), j(u′′ )) = inf ℓGadt (C ) : C ∈ ΩM (j(u′ ), j(u′′ ))

{

}

where ℓG (C ) denotes the length of the curve C with respect to the Riemannian metric G. Let c ∈ ΩΣ (u′ , u′′ ). As j is assumed to be an isometry (i.e. j∗ Gadt = GΣ ) one has dadt (j(u′ ), j(u′′ )) ≤ ℓGadt (j ◦ c) = ℓGΣ (c) and taking the infimum over ΩΣ (u′ , u′ ) yields dadt (j(u′ ), j(u′′ )) ≤ dΣ (u′ , u′′ ). (ii) The completion O+ (Σ ) is a metric space with the distance function dΣ (u , u ) = lim dΣ (uν , u′ν ) ′

ν→∞

where {uν }ν≥1 and {u′ν }ν≥1 are Cauchy sequences in (O+ (Σ ), dΣ ) representing u and u respectively. j induces the map ′

j : O+ (Σ ) → O+ (M, Σ )

adt

,

j(u) = lim j(uν ), ν→∞

where the limit is taken with respect to the dΣ and dadt metric topologies. Let u ∈ O+ (Σ ) and B ∈ O++ (1, 1). The action of O++ (1, 1) on O+ (Σ ) is given by u · B = lim RB (uν )

(90)

ν→∞

where RB : O+ (Σ ) → O+ (Σ ) is the right translation with B. Since

j(u · B) = j(u) · B˜ ,

B˜ ≡

(

B 0

0 I2

)

,

˜ for any u ∈ O+ (Σ ) it follows that [by (90)] j(u · B) = j(u) · B.



If β is a connection 1-form on a given principal bundle we denote by Γ (β ) = Ker(β ) the corresponding connection distribution. Lemma 10. Let Σ ⊂ M be a Lorentzian surface endowed with the sections (sT , s⊥ ). Then for any V ∈ Γ ψ T

(

j ωo(1,1) V = 0.

(∗

)

) (91)

Consequently Γ (ψ ) = Ker(j ωo(1,1) ). T



Proof. By the very definitions j is a right inverse to hT hT ◦ j = 1O(Σ ) .

(92)

Differentiating (92) one has (dj(p) hT )(dp j)Vp = Vp ∈ Γ (ψ T )p = (dj(p) hT )Γ (ω)j(p) for any p ∈ O(Σ ). Hence there is W ∈ Γ (ω)j(p) such that (dj(p) hT ) W − (dp j)Vp = 0.

[

]

(93)

Next let us apply ψ to both sides of (93) and use (87). It follows that T

( ) ωo(1,1) W − j∗ ωo(1,1) p V = 0. As W is ω-horizontal [by taking into account the direct sum decomposition o(1,)3) = o(1, 1)⊕o(2)⊕g] one has ωo(1,1) (W ) = 0 ( and we may conclude that (91) holds good. This yields Γ (ψ T )p ⊂ Ker j∗ ωo(1,1) p and the last statement in Lemma 10 follows by comparing dimensions. □ Let us observe that (again by the very definitions) h⊥ ◦ j = s ⊥ ◦ Π T .

(94)

Let a : [0, 1] → Σ be a C curve and let a : [0, 1] → O (Σ ) be a ψ -horizontal lift of a. Let A : [0, 1] → M be the curve given by A = ι ◦ a. Let AH : [0, 1] → O+ (M, Σ ) be a ω-horizontal lift of A. 1

H

+

T

164

E. Barletta et al. / Journal of Geometry and Physics 120 (2017) 142–168

Lemma 11. Let Σ be a Lorentzian surface in (M, Fθ ) with a flat normal connection (Dψ ⊥ = 0). Let sT : Σ → O(Σ ) be a given globally defined section. Let {Ni : i ∈ {1, 2}} be an orthonormal frame in T (Σ )⊥ such that ∇ ⊥ Ni = 0 for any i ∈ {1, 2} and let us consider the section s⊥ : Σ → O(T (Σ )⊥ ),

s⊥ (p)ei = Ni (p) , i ∈ {1, 2}.

Then, ( ) ∗ (i) s⊥ ψ ⊥ = 0. (ii) j : O(Σ ) → O(M, Σ ) is an isometry i.e. j∗ Gadt = GΣ . (iii) AH (t) = j(aH (t)) for any 0 ≤ t ≤ 1. adt (iv) j descends to a map of orbit spaces j : Σ → M and

˙ ⊂ ∂adt M. j Σ

( )

(95)

Proof. (i) Let p ∈ Σ and v ∈ Tp (Σ ). As v is tangent to Σ there is a C 1 curve p : (−ϵ, ϵ ) → Σ such that p(0) = p and p˙ (0) = v . Let us set b(t) = s⊥ (p(t)),

|t | < ϵ.

˙ is ψ ⊥ -horizontal. Then As Ni are parallel in the normal bundle the tangent vector b(t) Ker ψ ⊥

(

) b(t)

˙ = (dp(t) s⊥ )p˙ (t) ∋ b(t)

yields ⊥ 0 = ψb(t) (dp(t) s⊥ )p˙ (t) = (s⊥ )∗ ψ ⊥

[

] p(t)

p˙ (t)

and in particular for t = 0 (s⊥ )∗ ψ ⊥ v = 0.

[

]

(ii) Using (89), (92), (94) and (i) in Lemma 11, one has

( )∗

j ∗ ω = j ∗ hT

( )∗ ( )∗ ( )∗ ψ T + j∗ h⊥ ψ ⊥ = ψ T + Π T s⊥ ψ ⊥ = ψ T .

Also, for any u ∈ O(Σ ), (j∗ Θ )u = Θj(u) ◦ (du j) =

(( T )∗ h

ΘΣ

) j(u)

◦ (du j) = ΘΣ ,hT (j(u)) (dj(u) hT )(du j)

= ΘΣ , u . Finally,

j∗ Gadt = (j∗ ω) · (j∗ ω) + (j∗ Θ ) · (j∗ Θ ) = ψ T · ψ T + ΘΣ · ΘΣ = GΣ . (iii) Let u0 ∈ O+ (Σ ) be fixed and let aH and AH be respectively the horizontal lifts of a and A issuing at u0 and j(u0 ), i.e. aH = aH ( · ; u0 ) ,

AH = AH ( · ; j(u0 )) .

Let us also consider the curve C = j ◦ aH . Then both AH and C issue at j(u0 ) and AH is ω-horizontal. To see that C and AH coincide it suffices to check that C is ωhorizontal as well. Indeed, by (89),

ωC (t) C˙ (t) =

(( T )∗ h

ψT

) C (t)

C˙ (t) +

((

)∗

h⊥ ψ ⊥

) C (t)

C˙ (t)

= ψhTT (C (t)) (dC (t) hT )C˙ (t) + ψh⊥⊥ (C (t)) (dC (t) h⊥ )C˙ (t) and by (92), (dC (t) hT )C˙ (t) = (dC (t) hT )(daH (t) j)a˙ H (t) = a˙ H (t). Similarly, using (94), (dC (t) h⊥ )C˙ (t) = (dC (t) h⊥ )(daH (t) j)a˙ H (t) = (da(t) s⊥ )(daH (t) Π T )a˙ H (t)

= (da(t) s⊥ )a˙ (t). Therefore,

ωC (t) C˙ (t) = ψaTH (t) a˙ H (t) + ψs⊥⊥ (a(t)) (da(t) s⊥ )a˙ (t).

(96)

E. Barletta et al. / Journal of Geometry and Physics 120 (2017) 142–168

165

Now the first term in the right hand side of (96) vanishes because a˙ H (t) is ψ T -horizontal, while the second term vanishes by statement (i) in Lemma 11. Thus ωC (t) C˙ (t) = 0. This proves (iii). adt

(iv) The map j : O + (Σ ) → O+ (M, Σ ) is equivariant [by (ii) of Lemma 9] hence induces a map j : Σ → M . To see that j ˙ i.e. an orbit p = O++ (1, 1) · u maps the b-boundary of Σ into the adapted bundle boundary of M let us consider a point p ∈ Σ for some u ∈ O+ (Σ ) such that there is a C 1 curve a : [0, 1) → Σ with the following properties + and (1) limt →1− aH (t) = u in the ∫ 1O H(Σ ) −topology, (2) the improper integral 0 ∥a (t) 1 a˙ (t)∥ dt is convergent. Let A = ι ◦ a. We claim that lim AH (t) = j(u)

(97)

t →1−

in the O+ (M, Σ ) topology. Let {tν }ν≥1 ⊂ [0, 1) be a sequence such that tν → 1 as ν → ∞. Then, for any ϵ > 0, by (iii) in Lemma 11 and (ii) in Lemma 9, dadt AH (tν ) , j(u) = dadt j(aH (tν )) , j(u) ≤ dΣ aH (tν ) , u < ϵ

(

)

(

)

(

)

for any ν ≥ νϵ and some νϵ ≥ 1. Moreover, by (92), hT (AH (t)) = (hT ◦ j)aH (t) = aH (t). Therefore, the identity i j∗ Θ

(∗

) j(u)

) ( = (hT )∗ ΘΣ j(u) ,

u ∈ O(Σ ),

for u = aH (t) yields AH (t)−1 (dAH (t) Π ) = Θ(i◦j)(AH (t)) ◦ dAH (t) (i ◦ j)

( ) = i∗ j∗ Θ AH (t) = ΘΣ , hT (AH (t)) ◦ (dAH (t) hT ) ( )( ) = ΘΣ , aH (t) ◦ (dAH (t) hT ) = aH (t)−1 daH (t) Π T dAH (t) hT , which applied to A˙ H (t) gives

˙ = aH (t)−1 daH (t) Π T a˙ H (t) = aH (t)−1 a˙ (t). AH (t)−1 A(t)

(

)

Therefore 1



˙ ∥ dt = ∥AH (t)−1 A(t)



1

∥aH (t)−1 a˙ (t)∥ dt < ∞

0

0

implies j(p) ∈ ∂adt M.



7. Principal s1 -bundles over circles Γw0 To apply Theorem 5 to the total space Σ of a principal S 1 -subbundle over a circle Γw (with |f (w )| < µ) one needs to compute its normal curvature (and show that R⊥ = 0). This amounts to the calculation of dλ where λ ∈ Ω 1 (Σ ) is the differential 1-form determined by (99). Using Weingarten’s formula [cf. (98)] the calculation of the covariant derivative ∇ ⊥ N1 may be performed by computing the covariant derivative ∇ M N1 (with respect to the Levi-Civita connection of M) followed by projection on N2 , provided that explicit smooth extensions of Ni to tangent vector fields defined on the whole of π −1 (U) may be produced. This is indeed the case, as shown in the sequel. Let U ⊂ M be the open subset built in Section 4. The parameter ϕ − f (w ) may be thought of as the function

( π) φ : U → 0, , 2

φ (x) = ϕ (r , w) − f (w),

for any x = (r ei ϕ (r ,w) , w ) ∈ U. Let

⏐ ⏐ { } ∂0 W = (z , w) ∈ C2 : ⏐z − eif (w) ⏐ = 1, |f (w)| ≤ µ be the boundary of the worm domain with the ‘‘caps’’ removed. Note that A ⊂ ∂0 W . Let us set M0 = ∂0 W \ A so that U0 = (rei ϕ (r ,w) , w ) ∈ U : |f (w )| < µ

{

}

is an open subset in M0 . Note that Γw ⊂ U0 for any |f (w )| < µ. Let us consider the function g : U → R,

g(x) =

|w| cos φ (x)

, x ∈ U,

π2 (x) = w,

166

E. Barletta et al. / Journal of Geometry and Physics 120 (2017) 142–168

where π2 : C2 → C is the projection π2 (z , w ) = w . Moreover we shall need the functions hj : M → R, j ∈ {1, 2}, given by 1

w

e−i f (w) = h1 (x) + i h2 (x) ,

for any x ∈ M with π2 (x) = w . Let g π = g ◦ π and hπj = hj ◦ π : M → R be their vertical lifts. The upper script π will be omitted when convenient. We consider the vector fields Nj ∈ X(π −1 (U)) given by N1 = g π X ↑ + hπ1 S ,

(

N2 = g π Y ↑ − hπ2 S ,

)

(

)

so that point-wise restriction to Σ = π −1 (Γw0 ) gives the orthonormal frame {N1 , N2 } in the normal bundle T (Σ )⊥ considered in Lemma 8. Let ∇ M be the Levi-Civita connection of M and let us consider the differential 1-form λ ∈ Ω 1 (π −1 (U)) defined by

( ) λ(V ) = Fθ ∇ M V N1 , N2 for any V ∈ X(π −1 (U)). The same symbol λ denotes the pullback ι∗ λ ∈ Ω 1 (Σ ). We recall the Weingarten formula (cf. e.g. [23]) ⊥ ∇M V N = −AN V + ∇ V N

(98)

for any V ∈ X(Σ ) and any N ∈ C ∞ (T (Σ )⊥ ). Here AN and ∇ ⊥ are respectively the Weingarten, or shape, operator AN (associated to the normal section N) and the normal connection. The point-wise restriction of {N1 , N2 } to Σ is Fθ -orthonormal hence (by Weingarten’s formula)

∇ ⊥ N1 = λ ⊗ N2 ,

∇ ⊥ N2 = −λ ⊗ N1 .

(99)

In particular, if ⊥ ⊥ R⊥ (V , W ) = ∇ ⊥ V , ∇ W − ∇ [V , W ]

[

]

is the curvature of the normal connection then R⊥ (V , W )N1 = 2(dλ)(V , W )N2 ,

R⊥ (V , W )N2 = −2(dλ)(V , W )N1 .

We wish to compute the forms λ and dλ. To this end we recall that (cf. e.g. Lemma 2 in [16], p. 083504-26) W ↑ = (∇V W )↑ − (dθ )(V , W )T ↑ ∇M V↑ − [A(V , W ) + (dσ )(V ↑ , W ↑ )]S , M ↑ ∇ V ↑ T = (τ V + Φ V ) ↑ , V ↑ = (∇T V + Φ V )↑ + 2(dσ )(V ↑ , T ↑ )S , ∇M T↑

(100)

↑ ↑ S = ∇M ∇M S V = (JV ) , V↑

T ↑ = X↑ , ∇M T↑

∇M S S = 0,

↑ M ∇M S T = ∇ T ↑ S = 0,

for any V , W ∈ H(M), where ∇ is the Tanaka–Webster connection of (M , θ ). Also, Φ : H(M) → H(M) is given by Gθ (Φ V , W ) = (dσ )(V ↑ , W ↑ ) and X ∈ H(M) is determined by Gθ (X, V ) = 2(dσ )(T ↑ , V ↑ ). Lemma 12. The 1-form λ ∈ Ω 1 (π −1 (U)) is given by

λ(S) =

1 2

g 2 g11 ,

λ(X ↑ ) = g 2

{

λ(Y ) = −g ↑

2

1

}

h1 g11 + Gθ (∇X X , Y ) ,

2{

1 4

h2 g11 + Gθ (X , ∇Y Y )

}

λ(T ) = g Gθ (∇T X + Φ X , Y ), 2



Consequently

λ(Z ↑ ) =

1 2

g 2 g11

( h1 +

i 2

) h2

+ g 2 {Gθ (∇X X , Y ) + i Gθ (X , ∇Y Y )} .

(101)

E. Barletta et al. / Journal of Geometry and Physics 120 (2017) 142–168

167

Moreover, (101) yield (dλ)(S , T ↑ ) = − (dλ)(S , X ↑ ) = − (dλ)(S , Y ↑ ) = −

1 4 1 4 1 4

T g 2 g11 ,

(

)

X g 2 g11 ,

(

)

(102)

Y g 2 g11 ,

)

(

Consequently, (dλ)(S , Z ↑ ) = −

1 2

Z g 2 g11 .

(

)

Proof. The calculation of (dλ)(T ↑ , X ↑ ), (dλ)(T ↑ , Y ↑ ) and (dλ)(X ↑ , Y ↑ ) may be performed along the same lines, yet will not be required in the sequel. To prove Lemma 12 note that Gθ (X , Y ) = 0,

∥X ∥2 = ∥Y ∥2 =

1 2

g11 .

Moreover, (100) and JX = Y yield ↑ ↑ ∇M S X = Y ,

hence π ↑ π 2 2 λ(S) = Fθ (∇ M S N1 , N2 ) = g Fθ (Y , N2 ) = (g ) ∥Y ∥

leading to the first identity in (101). The proof of the remaining identities in (101) is similar and therefore omitted. Moreover, note that by (100) we have S , T ↑ = S , X ↑ = S , Y ↑ = 0.

[

]

[

]

[

]

Then, by (101), 2 (dλ)(S , T ↑ ) = S(λ(T ↑ )) − T ↑ (λ(S)) = −

1 2 1

2 (dλ)(S , X ↑ ) = S(λ(X ↑ )) − X ↑ (λ(S)) = − 2 (dλ)(S , Y ↑ ) = S(λ(Y ↑ )) − Y ↑ (λ(S)) = −

2 1 2

T (g 2 g11 ) ◦ π , X (g 2 g11 ) ◦ π, Y (g 2 g11 ) ◦ π.

This proves the lemma. □ Lemma 13. The open set U0 ⊂ M0 is foliated by the circles

{Γw : |f (w)| < µ}. If Γw is the leaf passing through the point x = (r ei ϕ (r ,w) , w ) ∈ U0 then x = γw (ϕ ) for the value ϕ = ϕ (r , w ) of the parameter. Then g11 (x) =

2 g(x)2

.

(103)

Proof. One has x = γw (ϕ ) if and only if ϕ = ϕ (r , w ). Moreover [by (42)] g11 (γw (ϕ )) =

2

|w|2

cos2 [ϕ − f (w )] =

2 g(γw (ϕ ))2

.

Lemma 14. For any |f (w )| < µ and any 0 < ϕ − f (w ) <

□ π 2

(dλ)γ ↑ (ϕ ; p0 ·ζ ) Sγ ↑ (ϕ ; p0 ·ζ ) , γ˙ (ϕ ; p0 · ζ ) = 0.

(



)

(104)

Consequently if Σ = π −1 (Γw ) then the normal connection of the immersion ι : Σ ↪→ M is flat (i.e. R⊥ = 0). Proof. Let p0 ∈ π −1 (U0 ) and ζ ∈ S 1 . To apply Lemma 12 one needs to extend γ˙ ↑ to a smooth vector field defined on the whole of π −1 (U0 ). A smooth extension is V = 2T − g 2 (h1 X − h2 Y ) ∈ X(π −1 (U0 )).

168

E. Barletta et al. / Journal of Geometry and Physics 120 (2017) 142–168

Indeed, by Lemma 7, ↑

γ˙ ↑ (ϕ ; p0 · ζ ) = Vγ ↑ (ϕ ; p

0 ·ζ )

because both sides above are σ -horizontal and project on γ˙ (ϕ ). On the other hand, by Lemma 12 and (103) in Lemma 13, 2(dλ)(S , V ↑ ) = −T g 2 g11 +

(

)

1 2

g 2 h1 X g 2 g11 − h2 Y g 2 g11

{

(

)

(

)}

= 0,

implying that ι∗ dλ = 0. □ Proof of Theorem 6. Let {N1 , N2 } be the Fθ -orthonormal frame in T (Σ )⊥ provided by Lemma 8. To produce a new frame νi = fij Nj which is parallel in the normal bundle (∇ ⊥ νi = 0) one solves the linear PDE’s system dfi1 = fi2 λ,

dfi2 = −fi1 λ,

(105)

whose integrability condition is dλ = 0 or equivalently R = 0. By Lemma 14 the real surface Σ = π (Γw ), |f (w )| < µ, has a flat normal connection. Let then {ν1 , ν2 } be a solution to (105) and let s⊥ be the corresponding section in O(T (Σ )⊥ ). Then (s⊥ )∗ ψ ⊥ = 0 by (i) in Lemma 11. The immersion j : O(Σ ) → O(M, Σ ) is then isometric [cf. (ii) in Lemma 11]. Then −1



adt

Lemma 9 applies so that j induces a map of completions j : O+ (Σ ) → O+ (M, Σ ) . According once again to Lemma 9 this is adt ˙ into the adapted equivariant hence gives rise to a map j : Σ → M . Finally [by (iv) in Lemma 11] j maps the b-boundary Σ bundle boundary ∂adt M. □ References [1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23]

K. Diederich, J.E. Fornaess, Pseudoconvex domains: an example with nontrivial Nebenhülle, Math. Ann. 225 (3) (1977) 275–292. S.G. Krantz, M.M. Peloso, Analysis and geometry on worm domains, J. Geom. Anal. 18 (2008) 478–510. M. Christ, Global C ∞ regularity of the ∂ -Neumann problem for worm domains, J. Amer. Math. Soc. 9 (4) (1996) 1171–1185. G.B. Folland, J.J. Kohn, The Neumann Problem for the Cauchy–Riemann Complex, Princeton University Press, Princeton, 1972. S.G. Krantz, Function Theory of Several Complex Variables, American Mathematical Society, Providence, 2001. S.M. Webster, Pseudohermitian structures on a real hypersurface, J. Differential Geom. 13 (1978) 25–41. C.J.S. Clarke, The Analysis of Space–Time Singularities, in: Cambridge Lecture Notes in Physics, Cambridge University Press, 2012. C. Fefferman, Monge–Ampère equations, He Bergman Kernel, and Geometry of Pseudoconvex Domains, Ann. of Math. (2) 103 (1976) 395–416; C. Fefferman, Monge–Ampère equations, He Bergman Kernel, and Geometry of Pseudoconvex Domains, Ann. of Math. (2) 104 (1976) 393–394. I. Naruki, On extendibility of isomorphisms of Cartan connections and biholomorphic mappings of bounded domains, Tohoku Math. J. 28 (1976) 117–122. C. Fefferman, The Bergman kernel and biholomorphic maps of pseudoconvex domains, Invent. Math. 26 (1975) 1–65. S.-Y. Cheng, S.-T. Yau, On the existence of a complete Kähler metric on non-compact complex manifolds and the regularity of Fefferman’s equation, Comm. Pure Appl. Math. 33 (1980) 507–544. S-Y. Li, On the existence and regularity of Dirichlet problem for complex Monge–Ampère equations on weakly pseudoconvex domains, Calc. Var. 20 (2004) 119–132. J.M. Lee, The Fefferman metric and pseudohermitian invariants, Trans. Amer. Math. Soc. 296 (1) (1986) 411–429. J.A. Thorpe, Curvature invariants and space–time singularities, J. Math. Phys. 18 (5) (1977) 960–964. B.G. Schmidt, A new definition of singular points in general relativity, Gen. Relativity Gravitation 1 (3) (1971) 269–280. E. Barletta, S. Dragomir, H. Urakawa, Yang–Mills fields on CR manifolds, J. Math. Phys. 47 (8) (2006) 083504 1–41. R.A. Johnson, The bundle boundary in some special cases, J. Math. Phys. 18 (1977) 898. B. Bosshard, On the b-boundary of the closed Friedman model, Comm. Math. Phys. 46 (1976) 263–268. S. Dragomir, G. Tomassini, Differential Geometry and Analysis on CR Manifolds, in: Progress in Mathematics, vol. 246, Birkhäuser, Boston-Basel-Berlin, 2006. C.R. Graham, On Sparling’s characterization of Fefferman metrics, Amer. J. Math. 109 (1987) 853–874. S. Kobayashi, K. Nomizu, Foundations of Differential Geometry, Vol. 1, Interscience Publishers, New York, 1963 vol. 2, 1969. H. Friedrich, Construction and properties of space–time b-boundaries, Gen. Relativity Gravitation 5 (1974) 681–697. B.-Y. Chen, Geometry of submanifolds, in: Pure and Appl. Math., Marcel Dekker, Inc., New York, 1973.