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

The Newlander-Nirenberg theorem for
complex bb-manifolds

Tatyana Barron and Michael Francis
(October 12, 2023)
Abstract

Melrose defined the bb-tangent bundle of a smooth manifold MM with boundary as the vector bundle whose sections are vector fields on MM tangent to the boundary. Mendoza defined a complex bb-manifold as a manifold with boundary together with an involutive splitting of the complexified bb-tangent bundle into complex conjugate factors. In this article, we prove complex bb-manifolds have a single local model depending only on dimension. This can be thought of as the Newlander-Nirenberg theorem for complex bb-manifolds: there are no “local invariants”. Our proof uses Mendoza’s existing result that complex bb-manifolds do not have “formal local invariants” and a singular coordinate change trick to leverage the classical Newlander-Nirenberg theorem.

1 Introduction

Throughout, we conflate 2=\mathbb{R}^{2}=\mathbb{C} and use coordinates z=(x,y)=x+iyz=(x,y)=x+iy interchangeably. The word “smooth” means “infinitely real-differentiable”, not “holomorphic”.

Recall that a complex manifold may equivalently defined either by a holomorphic atlas or by an integrable almost-complex structure. The equivalence of these two definitions is given by the Newlander-Nirenberg theorem [8]. In more detail, a complex structure on a smooth manifold MM is an involutive subbundle T0,1MT^{0,1}M of the complexified tangent bundle such that TM=T1,0MT0,1M\mathbb{C}TM=T^{1,0}M\oplus T^{0,1}M where T1,0M:-T0,1M¯T^{1,0}M\coloneq\overline{T^{0,1}M}. The Newlander-Nirenberg theorem implies that every point belongs to a coordinate chart (x1,y1,,xn,yn)=(z1,,zn)(x_{1},y_{1},\ldots,x_{n},y_{n})=(z_{1},\ldots,z_{n}) in which T0,1MT^{0,1}M is spanned by

z¯j:-12(xj+iyj)\displaystyle\partial_{\overline{z}_{j}}\coloneq\tfrac{1}{2}(\partial_{x_{j}}+i\partial_{y_{j}}) j=1,,n.\displaystyle j=1,\ldots,n.

Accordingly, T1,0MT^{1,0}M is spanned by

zj:-12(xjiyj)\displaystyle\partial_{z_{j}}\coloneq\tfrac{1}{2}(\partial_{x_{j}}-i\partial_{y_{j}}) j=1,,n.\displaystyle j=1,\ldots,n.

In this article, a bb-manifold is a smooth manifold MM equipped with a closed hypersurface ZMZ\subseteq M referred to as the singular locus. The bb-tangent bundle is the vector bundle TbM{{}^{b}TM} over MM whose smooth sections are the vector fields on MM that are tangent along ZZ. These terminologies stem from the bb-geometry of Melrose [6], also called log-geometry.

Mendoza [7] defined a complex bb-structure on a bb-manifold MM to be an involutive subbundle T0,1bM{{}^{b}T^{0,1}}M of the complexified bb-tangent bundle such that TbM=T1,0bMT0,1bM\mathbb{C}{{}^{b}TM}={{}^{b}T^{1,0}}M\oplus{{}^{b}T^{0,1}}M, where T1,0bM:-T0,1bM¯{{}^{b}T^{1,0}}M\coloneq\overline{{{}^{b}T^{0,1}}M}. Thus, complex bb-structures are defined exactly analogously to complex structures, replacing the tangent bundle by the bb-tangent bundle.

To complete the analogy, one would like there to be a Newlander-Nirenberg type result to the effect that all complex bb-structures are locally isomorphic to a standard model depending only dimension. Such a result would empower one to give a second (equivalent) definition of a complex bb-manifold in terms of an appropriately defined bb-holomorphic atlas. A logical candidate model for a complex bb-manifolds is M=2n+2=n+1M=\mathbb{R}^{2n+2}=\mathbb{C}^{n+1}, n0n\geq 0 equipped with coordinates (x0,y0,,xn,yn)=(z0,,zn)(x_{0},y_{0},\ldots,x_{n},y_{n})=(z_{0},\ldots,z_{n}) with T0,1bM{{}^{b}T^{0,1}}M spanned by

z¯0b\displaystyle{{}^{b}\partial}_{\overline{z}_{0}} :-12(x0x0+iy0)\displaystyle\coloneq\tfrac{1}{2}(x_{0}\partial_{x_{0}}+i\partial_{y_{0}})
z¯j\displaystyle\partial_{\overline{z}_{j}} :-12(xj+iyj)\displaystyle\coloneq\tfrac{1}{2}(\partial_{x_{j}}+i\partial_{y_{j}}) j=1,,n.\displaystyle j=1,\ldots,n.

Accordingly, T1,0bM{{}^{b}T^{1,0}}M is spanned by

z0b\displaystyle{{}^{b}\partial}_{z_{0}} :-12(x0x0iy0)\displaystyle\coloneq\tfrac{1}{2}(x_{0}\partial_{x_{0}}-i\partial_{y_{0}})
zj\displaystyle\partial_{z_{j}} :-12(xjiyj)\displaystyle\coloneq\tfrac{1}{2}(\partial_{x_{j}}-i\partial_{y_{j}}) j=1,,n.\displaystyle j=1,\ldots,n.

Our goal in this article is to show that, locally near points on the singular locus, every complex bb-manifold of dimension 2n+22n+2 is indeed isomorphic to the model example above. Mendoza [7] already took up this problem and partially settled it, up to deformation by terms whose Taylor expansions vanish along the singular locus. Indeed, Mendoza’s result, which we state below, will play a crucial role.

Theorem 1.1 ([7], Proposition 5.1).

With n0n\geq 0, let MM be a (2n+2)(2n+2)-dimensional complex bb-manifold with singular locus ZZ. Then, every pZp\in Z has an open neighbourhood UU on which there are smooth local coordinates (x0,y0,,xn,yn)=(z0,,zn)(x_{0},y_{0},\ldots,x_{n},y_{n})=(z_{0},\ldots,z_{n}) centered at pp with x0x_{0} vanishing on ZZ and smooth, complex vector fields Γ0,,Γn\Gamma_{0},\ldots,\Gamma_{n} on UU vanishing to infinite order on ZZ such that

L0\displaystyle L_{0} :-z¯0b+Γ0\displaystyle\coloneq{{}^{b}\partial}_{\overline{z}_{0}}+\Gamma_{0}
Lj\displaystyle L_{j} :-z¯j+Γj\displaystyle\coloneq\partial_{\overline{z}_{j}}+\Gamma_{j} j=1,,n\displaystyle j=1,\ldots,n

is a frame for T0,1bM{{}^{b}T^{0,1}}M over UU. Moreover, each of Γ0,,Γn\Gamma_{0},\ldots,\Gamma_{n} is a linear combination of z0b,z1,,zn{{}^{b}\partial}_{z_{0}},\partial_{z_{1}},\ldots,\partial_{z_{n}} whose coefficients are smooth, complex-valued functions on UU vanishing to infinite order on ZZ.

The contribution of the present article will be to improve the above result by showing that the deformation terms Γ0,,Γn\Gamma_{0},\ldots,\Gamma_{n} can be got rid of. Thus, our main result is the following.

Theorem 1.2.

With n0n\geq 0, let MM be a (2n+2)(2n+2)-dimensional complex bb-manifold with singular locus ZZ. Then, every pZp\in Z has an open neighbourhood UU on which there are smooth local coordinates (x0,y0,,xn,yn)=(z0,,zn)(x_{0},y_{0},\ldots,x_{n},y_{n})=(z_{0},\ldots,z_{n}) centered at pp with x0x_{0} vanishing on ZZ such that

z¯0b\displaystyle{{}^{b}\partial}_{\overline{z}_{0}} :-12(x0x0+iy0)\displaystyle\coloneq\tfrac{1}{2}(x_{0}\partial_{x_{0}}+i\partial_{y_{0}})
z¯j\displaystyle\partial_{\overline{z}_{j}} :-12(xj+iyj)\displaystyle\coloneq\tfrac{1}{2}(\partial_{x_{j}}+i\partial_{y_{j}}) j=1,,n\displaystyle j=1,\ldots,n

is a frame for T0,1bM{{}^{b}T^{0,1}}M over UU.

Readers familiar with the techniques needed to prove the classical Newlander-Nirenberg theorem will be unsurprised that nonellipticity of the operator z¯b=12(xx+iy){{}^{b}\partial}_{\overline{z}}=\frac{1}{2}(x\partial_{x}+i\partial_{y}) is at the heart of the matter. The difficulty is that, without ellipticity, we do not automatically have good existence theory for solutions to deformations of this operator. The main insight of this article is that it is possible to get around this difficulty by leveraging the polar coordinate change g:22g:\mathbb{R}^{2}\to\mathbb{R}^{2}, g(x,y)=xeiyg(x,y)=xe^{iy} (negative radii permitted). Notice that z¯bg=0{{}^{b}\partial}_{\overline{z}}g=0, i.e. gg is “bb-holomorphic”.

Let us summarize the contents of this paper. In Section 2, we collect definitions and basic results on complex bb-manifolds. In Section 3, we prove in the two-dimensional case that nontrivial bb-holomorphic functions exist locally. In Section 4, we use the latter existence result to prove the main result Theorem 1.2 in the two-dimensional case. Sections 5 and 6 extend the results of Sections 3 and 4 to the general case. Finally, in Section 7, we situate the singular coordinate change g(x,y)=xeiyg(x,y)=xe^{iy} in a more general context which will be relevant for planned future work on complex bkb^{k}-manifolds (see [1] for the definition of a complex bkb^{k}-manifold).

2 Background

The language of bb-geometry, also known as log-geometry, was introduced by Melrose [6]. One encounters two superficially different formulations of bb-geometry in the literature. In the original approach, one works on a manifold with boundary. Other authors instead use a manifold without boundary that is equipped with a given hypersurface [2]. We follow the latter approach.

Definition 2.1.

A 𝐛\mathbf{b}-manifold is a smooth manifold MM together with a closed hypersurface ZMZ\subseteq M which we refer to as the singular locus. An isomorphism of bb-manifolds is a diffeomorphism that preserves the singular loci.

Note ordinary smooth manifolds may be considered as bb-manifolds, taking Z=Z=\varnothing.

Definition 2.2.

A 𝐛\mathbf{b}-vector field on a bb-manifold M=(M,Z)M=(M,Z) is a smooth vector field on MM that is tangent to ZZ. The collection of bb-vector fields on MM is denoted 𝔛b(M)𝔛(M){{}^{b}\mathfrak{X}}(M)\subseteq\mathfrak{X}(M).

Example 2.3.

Consider the bb-manifold M=n+1M=\mathbb{R}^{n+1} with coordinates (x0,x1,,xn)(x_{0},x_{1},\ldots,x_{n}) and singular locus Z={0}×nZ=\{0\}\times\mathbb{R}^{n}. Then, 𝔛b(M)=x0x0,x1,,xn{{}^{b}\mathfrak{X}}(M)=\langle x_{0}\partial_{x_{0}},\partial_{x_{1}},\ldots,\partial_{x_{n}}\rangle, the free C(M)C^{\infty}(M)-module generated by x0x0,x1,,xnx_{0}\partial_{x_{0}},\partial_{x_{1}},\ldots,\partial_{x_{n}}.

The example above completely captures the local structure of bb-manifolds. Accordingly, for any bb-manifold MM, 𝔛b(M){{}^{b}\mathfrak{X}}(M) is a projective C(M)C^{\infty}(M)-module closed under Lie bracket. Applying Serre-Swan duality to the inclusion 𝔛b(M)𝔛(M){{}^{b}\mathfrak{X}}(M)\to\mathfrak{X}(M), one has a corresponding Lie algebroid TbM{{}^{b}TM} whose anchor map, denoted ρ:TbMTM\rho:{{}^{b}TM}\to TM, induces (abusing notation) an identification C(M;TbM)=𝔛b(M)C^{\infty}(M;{{}^{b}TM})={{}^{b}\mathfrak{X}}(M).

Definition 2.4.

For a bb-manifold MM, the Lie algebroid TbM=(TbM,ρ,[,]){{}^{b}TM}=({{}^{b}TM},\rho,[\cdot,\cdot]) satisfying C(M;TbM)=𝔛b(M)C^{\infty}(M;{{}^{b}TM})={{}^{b}\mathfrak{X}}(M) described above is called the 𝐛\mathbf{b}-tangent bundle of MM.

In Example 2.3, x0x0,x1,,xnx_{0}\partial_{x_{0}},\partial_{x_{1}},\ldots,\partial_{x_{n}} is a global frame for TbM{{}^{b}TM}. The anchor map ρ\rho descends from evaluation of vector fields. Thus, over MZM\setminus Z, ρ\rho is an isomorphism and, over ZZ, the kernel of ρ\rho is the line bundle spanned by x0x0x_{0}\partial_{x_{0}}. In general, for any bb-manifold MM, the bundles TbM{{}^{b}TM} and TMTM are canonically isomorphic over MZM\setminus Z via ρ\rho.

If θ:M1M2\theta:M_{1}\to M_{2} is an isomorphism of bb-manifolds, it is clear that θ\theta preserves the associated modules of bb-vector fields. By Serre-Swan duality, θ\theta induces a Lie algebroid isomorphism

θb:TbM1TbM2.\displaystyle{{}^{b}\theta}_{*}:{{}^{b}TM_{1}}\to{{}^{b}TM_{2}}. (2.1)

Over MiZiM_{i}\setminus Z_{i}, if we apply the aforementioned natural identifications of the bb-tangent bundle and usual tangent bundle, θb{{}^{b}\theta_{*}} coincides with θ:TM1TM2\theta_{*}:TM_{1}\to TM_{2}, the usual induced isomorphism given by pushforward of tangent vectors.

Definition 2.5.

The 𝐛\mathbf{b}-cotangent bundle TbM{{}^{b}TM^{*}} of a bb-manifold MM is the dual of the bb-tangent bundle TbM{{}^{b}TM}.

The dual ρ:TMTbM\rho^{*}:TM^{*}\to{{}^{b}TM^{*}} of the anchor map ρ:TbMTM\rho:{{}^{b}TM}\to TM is likewise an isomorphism when restricted over MZM\setminus Z. In particular, ρ\rho^{*} is injective on a dense open set, and so (what is equivalent) induces an injective mapping C(M;TM)C(M;TbM)C^{\infty}(M;TM^{*})\to C^{\infty}(M;{{}^{b}TM^{*}}). Put in simpler terms, a 1-form ωC(M;TM)\omega\in C^{\infty}(M;TM^{*}) is fully determined by its pairing with bb-vector fields. The 𝐛\mathbf{b}-exterior derivative of a smooth function ff on MM is defined as:

dbf:-ρ(df).\displaystyle{{}^{b}d}f\coloneq\rho^{*}(df). (2.2)

Actually, one may quite legitimately regard dfdf and dbf{{}^{b}d}f as identical, with the latter notation merely hinting that one only intends to pair the form with bb-vector fields.

A general philosophy of bb-calculus is that many classical geometries admit bb-analogues in which the role of tangent bundle is played by the bb-tangent bundle. A notable example is bb-symplectic geometry ([4], [5]). Complex bb-geometry was introduced by Mendoza [7] who furthermore made a systematic study of the bb-Dolbeault complex. We remark that the basic idea of a complex bb-structure is mentioned in passing on pp. 218 of [6].

Definition 2.6.

A complex 𝐛\mathbf{b}-structure on an (even-dimensional) bb-manifold MM is a complex subbundle T0,1bM{{}^{b}T^{0,1}}M of the complexified bb-tangent bundle satisfying:

  1. (i)

    TbM=T1,0bMT0,1bM\mathbb{C}{{}^{b}TM}={{}^{b}T^{1,0}}M\oplus{{}^{b}T^{0,1}}M, where T1,0bM:-T0,1bM¯{{}^{b}T^{1,0}}M\coloneq\overline{{{}^{b}T^{0,1}}M},

  2. (ii)

    T0,1bM{{}^{b}T^{0,1}}M is involutive.

A complex 𝐛\mathbf{b}-manifold is a bb-manifold equipped with a complex bb-structure. An isomorphism θ:M1M2\theta:M_{1}\to M_{2} of complex bb-manifolds is an isomorphism of their underlying bb-manifolds satisfying θb(T0,1bM1)=T0,1bM2{{}^{b}\theta_{*}}({{}^{b}T^{0,1}}M_{1})={{}^{b}T^{0,1}}M_{2}. Here, by abuse of notation, θb{{}^{b}\theta_{*}} denotes the complexification of the isomorphism θb:TbM1TbM2{{}^{b}\theta_{*}}:{{}^{b}TM_{1}}\to{{}^{b}TM_{2}} defined above (2.1).

Because TbM{{}^{b}TM} is naturally isomorphic to TMTM away from ZZ, a bb-complex structure on MM in particular gives a complex structure in the usual sense on MZM\setminus Z. In particular, one may consider ordinary complex manifolds as special cases of complex bb-manifolds with Z=Z=\varnothing.

Actually, for any complex bb-manifold MM, the restricted complex structure on MZM\setminus Z determines the bb-complex structure, as the following proposition shows. Of course, not all complex structures on MZM\setminus Z are the restriction of a (unique) bb-complex structure on MM, just those that degenerate in a particular way at ZZ.

Proposition 2.7.

Let MiM_{i} be a complex bb-manifold with singular locus ZiZ_{i} for i=1,2i=1,2. Suppose that θ:M1M2\theta:M_{1}\to M_{2} is a diffeomorphism such that:

  1. (i)

    θ(Z1)=Z2\theta(Z_{1})=Z_{2} (i.e. θ\theta is an isomorphism of bb-manifolds),

  2. (ii)

    θ\theta restricts to an isomorphism of (ordinary) complex manifolds M1Z1M2Z2M_{1}\setminus Z_{1}\to M_{2}\setminus Z_{2}.

Then, θ\theta is an isomorphism of complex bb-manifolds.

Proof.

From (i), we have an induced isomorphism on the complexified bb-tangent bundles θb:TbM1TbM2{{}^{b}\theta_{*}}:\mathbb{C}{{}^{b}T}M_{1}\to\mathbb{C}{{}^{b}TM_{2}}. From (ii), and applying the natural isomorphisms of bb-tangent bundle and ordinary tangent bundle away from the singular loci, the subbundles θb(T0,1bM1){{}^{b}\theta_{*}}({{}^{b}T^{0,1}}M_{1}) and T0,1bM2{{}^{b}T^{0,1}}M_{2} of TbM2\mathbb{C}{{}^{b}TM_{2}} agree away from Z2Z_{2}. The conclusion follows by a continuity argument. ∎

Taking duals, the splitting TbM=T1,0bMT0,1bM\mathbb{C}{{}^{b}TM}={{}^{b}T^{1,0}}M\oplus{{}^{b}T^{0,1}}M given by a complex bb-structure induces an associated splitting of the complexified bb-cotangent bundle TbM\mathbb{C}{{}^{b}TM^{*}} and of the complexified bb-exterior derivative db:C(M,)C(M;TbM){{}^{b}d}:C^{\infty}(M,\mathbb{C})\to C^{\infty}(M;\mathbb{C}{{}^{b}TM^{*}}), as tabulated below:

TbM=T1,0bMT0,1bM\displaystyle\mathbb{C}{{}^{b}TM^{*}}={{}^{b}T^{1,0}}M^{*}\oplus{{}^{b}T^{0,1}}M^{*}
db=b+¯b\displaystyle{{}^{b}d}={{}^{b}\partial}+{{}^{b}{\overline{\partial}}}
b:C(M,)C(M;T1,0bM)\displaystyle{{}^{b}\partial}:C^{\infty}(M,\mathbb{C})\to C^{\infty}(M;{{}^{b}T^{1,0}}M^{*})
¯b:C(M,)C(M;T0,1bM)\displaystyle{{}^{b}{\overline{\partial}}}:C^{\infty}(M,\mathbb{C})\to C^{\infty}(M;{{}^{b}T^{0,1}}M^{*})

In effect, bf{{}^{b}\partial}f, respectively ¯bf{{}^{b}{\overline{\partial}}}f, is dfdf restricted to the bb-(1,0)(1,0)-vector fields, respectively the bb-(0,1)(0,1)-vector fields (that is to say, sections of T1,0bM{{}^{b}T^{1,0}}M, respectively sections of T0,1bM{{}^{b}T^{0,1}}M).

Definition 2.8.

A bb-holomorphic function on a complex bb-manifold MM (or an open subset thereof) is a smooth function f:Mf:M\to\mathbb{C} satisfying ¯bf=0{{}^{b}{\overline{\partial}}}f=0. We write 𝒪b(M){{}^{b}\mathcal{O}}(M) for the collection of bb-holomorphic functions on MM.

In more concrete terms, ff is bb-holomorphic if Xf=0Xf=0 for every bb-(0,1)(0,1)-vector field XX or (what is sufficient), all XX in a given frame for T0,1bM{{}^{b}T^{0,1}}M. Because T0,1bM{{}^{b}T^{0,1}}M is involutive, 𝒪b(M){{}^{b}\mathcal{O}}(M) is a ring with respect to pointwise-multiplication.

Example 2.9.

Consider the bb-manifold M=2=M=\mathbb{R}^{2}=\mathbb{C} with singular locus Z={0}×Z=\{0\}\times\mathbb{R}. Then, 𝔛b(M)=xx,y{{}^{b}\mathfrak{X}}(M)=\langle x\partial_{x},\partial_{y}\rangle, the free C(2)C^{\infty}(\mathbb{R}^{2})-module generated by xxx\partial_{x} and y\partial_{y}. An example of a complex bb-structure for MM has T0,1bM{{}^{b}T^{0,1}}M spanned by

z¯b:-12(xx+iy).\displaystyle{{}^{b}\partial}_{\overline{z}}\coloneq\tfrac{1}{2}(x\partial_{x}+i\partial_{y}).

and, correspondingly, T1,0bM{{}^{b}T^{1,0}}M spanned by

zb:-12(xxiy).\displaystyle{{}^{b}\partial}_{z}\coloneq\tfrac{1}{2}(x\partial_{x}-i\partial_{y}).

A function ff is bb-holomorphic precisely when z¯bf=0{{}^{b}\partial}_{\overline{z}}f=0. An important globally-defined bb-holomorphic function on MM was mentioned in the introduction

g:2\displaystyle g:\mathbb{R}^{2}\to\mathbb{C} g(x,y)=xeiy.\displaystyle g(x,y)=xe^{iy}.

More generally, if hh is a usual holomorphic function defined near 00\in\mathbb{C}, then f=hgf=h\circ g is a bb-holomorphic function defined near 0M0\in M. There also exist bb-holomorphic functions defined near 0 not of the latter form and, indeed, not real-analytic near 0. For example, on the domain π4<y<π4-\frac{\pi}{4}<y<\frac{\pi}{4}, one has a bb-holomorphic ff defined by f(x,y)=exp(1g(x,y))f(x,y)=\exp(\frac{-1}{g(x,y)}) for x>0x>0 and f(x,y)=0f(x,y)=0 for x0x\leq 0. These examples are also noted in [1], Example 8.1.

Remark 2.10.

Observe that the bb-holomorphic function g(x,y)=xeiyg(x,y)=xe^{iy} in the above example may be conceptualized as g(x,y)=ϕiy(x)g(x,y)=\phi_{iy}(x), where ϕt(x)=etx\phi_{t}(x)=e^{t}x is the flow of xxx\partial_{x}. This somewhat cryptic remark will be expanded on in Section 7.

Example 2.11.

Consider the bb-manifold M=2n+2=n+1M=\mathbb{R}^{2n+2}=\mathbb{C}^{n+1} with Z={0}×2n+1Z=\{0\}\times\mathbb{R}^{2n+1} as its singular locus. Introduce coordinates (x0,y0,,xn,yn)=(z0,,zn)(x_{0},y_{0},\ldots,x_{n},y_{n})=(z_{0},\ldots,z_{n}). Thus,

𝔛b(M)=x0x0,y0,x1,y1,,xn,yn.\displaystyle{{}^{b}\mathfrak{X}}(M)=\langle x_{0}\partial_{x_{0}},\partial_{y_{0}},\partial_{x_{1}},\partial_{y_{1}},\ldots,\partial_{x_{n}},\partial_{y_{n}}\rangle.

An example of a complex bb-structure T0,1bM{{}^{b}T^{0,1}}M for MM is the one spanned by:

z¯0b\displaystyle{{}^{b}\partial}_{\overline{z}_{0}} :-12(x0x0+iy0)\displaystyle\coloneq\tfrac{1}{2}(x_{0}\partial_{x_{0}}+i\partial_{y_{0}})
z¯j\displaystyle\partial_{\overline{z}_{j}} :-12(xj+iyj)\displaystyle\coloneq\tfrac{1}{2}(\partial_{x_{j}}+i\partial_{y_{j}}) j=1,,n.\displaystyle j=1,\ldots,n.

Correspondingly, a frame for T1,0bM{{}^{b}T^{1,0}}M is:

z0b\displaystyle{{}^{b}\partial}_{z_{0}} :-12(x0x0iy0)\displaystyle\coloneq\tfrac{1}{2}(x_{0}\partial_{x_{0}}-i\partial_{y_{0}})
zj\displaystyle\partial_{z_{j}} :-12(xjiyj)\displaystyle\coloneq\tfrac{1}{2}(\partial_{x_{j}}-i\partial_{y_{j}}) j=1,,n.\displaystyle j=1,\ldots,n.

The function g(x0,y0,,xn,yn)=x0eiy0g(x_{0},y_{0},\ldots,x_{n},y_{n})=x_{0}e^{iy_{0}} from Example 2.9 is still bb-holomorphic, as are the complex coordinate functions zj=xj+iyjz_{j}=x_{j}+iy_{j}, j=1,,nj=1,\ldots,n. Additional bb-holomorphic functions may be obtained by applying an (n+1)(n+1)-variable holomorphic function (in the usual sense) to g,z1,,zng,z_{1},\ldots,z_{n}.

Example 2.12.

Consider the complex bb-manifold N=2N=\mathbb{R}^{2} with singular locus Z:-{0}×Z\coloneq\{0\}\times\mathbb{R} and with T0,1bN{{}^{b}T^{0,1}}N spanned by

L:-12(xx+i(xyx+(1+y2)y)).L\coloneq\tfrac{1}{2}\Big{(}x\partial_{x}+i(-xy\partial_{x}+(1+y^{2})\partial_{y})\Big{)}.

This complex bb-manifold NN is isomorphic to a neighbourhood of the origin of the complex bb-manifold M=2M=\mathbb{R}^{2} with T0,1bM{{}^{b}T^{0,1}}M spanned by z¯b:-12(xx+iy){{}^{b}\partial}_{\overline{z}}\coloneq\frac{1}{2}(x\partial_{x}+i\partial_{y}) from Example 2.9. Indeed, the diffeomorphism G:×(π2,π2)2G:\mathbb{R}\times(-\frac{\pi}{2},\frac{\pi}{2})\to\mathbb{R}^{2} defined by G(x,y)=(xcosy,tany)G(x,y)=(x\cos y,\tan y) satisfies G(z¯b)=LG_{*}({{}^{b}\partial}_{\overline{z}})=L. Thus, bb-holomorphic functions on NN may be obtained by pushforward through GG. For example, the pushforward of g(x,y)=xeiyg(x,y)=xe^{iy} by GG is p(x,y)=x+ixyp(x,y)=x+ixy which indeed satisfies Lp=0Lp=0.

For a general complex bb-manifold MM, it is not immediately clear whether, locally near points on the singular locus ZZ, any nonconstant bb-holomorphic functions must exist at all. This article will confirm that they do (Sections 3 and 5).

Note that a bb-holomorphic function on a complex bb-manifold MM restricts to a holomorphic function in the usual sense on the complex manifold MZM\setminus Z. Indeed, in a similar spirit to Proposition 2.7, if ff is smooth and restricts to a holomorphic function on MZM\setminus Z, then it is bb-holomorphic on MM by a continuity argument.

Recall that, in the case of ordinary complex manifolds (defined using integrable almost-complex structures), the existence of local holomorphic functions is closely-tied to the existence of charts. The proposition below, which will be used in Sections 4 and 6, illustrates this principle. We include its (standard) proof for the sake of completeness.

Proposition 2.13.

Let MM be a complex manifold. Let ΩM\Omega\subseteq M and Ωn\Omega^{\prime}\subseteq\mathbb{C}^{n} be open. Suppose f1,,fn:Ωf_{1},\ldots,f_{n}:\Omega\to\mathbb{C} are holomorphic functions such that θ:ΩΩ\theta:\Omega\to\Omega^{\prime}, θ(p)=(f1(p),,fn(p))\theta(p)=(f_{1}(p),\ldots,f_{n}(p)) defines a diffeomorphism. Then, θ\theta_{*} maps T0,1MT^{0,1}M onto T0,1nT^{0,1}\mathbb{C}^{n} over Ω\Omega where, by definition, the latter is spanned by z¯j:-12(xj+iyj)\partial_{\overline{z}_{j}}\coloneq\frac{1}{2}(\partial_{x_{j}}+i\partial_{y_{j}}), j=1,,nj=1,\ldots,n.

Proof.

Let LC(Ω;T0,1M)L\in C^{\infty}(\Omega;T^{0,1}M) be a (0,1)(0,1)-vector field defined on Ω\Omega. Using Lfj=0Lf_{j}=0 for j=1,,nj=1,\ldots,n and θ(fj)=zj\theta_{*}(f_{j})=z_{j} (the jjth coordinate function) for j=1,,nj=1,\ldots,n, we obtain

θ(L)zj=0.\theta_{*}(L)z_{j}=0.

If we write θ(L)=j=1nαjx+βjy\theta_{*}(L)=\sum_{j=1}^{n}\alpha_{j}\partial_{x}+\beta_{j}\partial_{y}, where αj,βj\alpha_{j},\beta_{j} are smooth, complex-valued functions on \mathbb{C}, then the above gives αj+iβj=0\alpha_{j}+i\beta_{j}=0 for j=1,,nj=1,\ldots,n. Thus, we obtain

θ(L)=j=12iβjz¯j,\theta_{*}(L)=\sum_{j=1}-2i\beta_{j}\partial_{\overline{z}_{j}},

so that θ(L)\theta_{*}(L) is a (0,1)(0,1) vector field on Ω\Omega^{\prime}. ∎

In the case of complex bb-manifolds, bb-holomorphic functions cannot directly play the role of coordinate functions of charts. To see this, note that any solution f:2f:\mathbb{R}^{2}\to\mathbb{C} of z¯bf=0{{}^{b}\partial}_{\overline{z}}f=0 (where z¯b:-12(xx+iy){{}^{b}\partial}_{\overline{z}}\coloneq\frac{1}{2}(x\partial_{x}+i\partial y)) is necessarily constant along Z={0}×Z=\{0\}\times\mathbb{R}.

3 Existence of bb-holomorphic functions: dimension 22

In this section, we show that, near any point on the singular locus of a two-dimensional complex bb-manifold, there is a nontrivial bb-holomorphic function. Here, “nontrivial” means the derivative normal to the singular locus is nonzero. From Theorem 1.1, this amounts to the following (as usual zb:-12(xxiy){{}^{b}\partial}_{z}\coloneq\frac{1}{2}(x\partial_{x}-i\partial_{y}), z¯b:-12(xx+iy){{}^{b}\partial}_{\overline{z}}\coloneq\frac{1}{2}(x\partial_{x}+i\partial_{y})).

Theorem 3.1.

Consider a complex bb-manifold M=2M=\mathbb{R}^{2} with singular locus Z={0}×Z=\{0\}\times\mathbb{R} and with T0,1bM{{}^{b}T^{0,1}}M spanned by

L:-z¯b+γzb\displaystyle L\coloneq{{}^{b}\partial}_{\overline{z}}+\gamma\cdot{{}^{b}\partial}_{z}

where γ:2\gamma:\mathbb{R}^{2}\to\mathbb{C} is a smooth function vanishing to infinite order on ZZ. Then, there is a neighbourhood UMU\subseteq M of (0,0)(0,0) and a bb-holomorphic function f:Uf:U\to\mathbb{C} that vanishes on ZZ and satisfies xf(0,0)=1\partial_{x}f(0,0)=1.

The proof of Theorem 3.1 will rely on the following three elementary lemmas whose proofs we omit.

Firstly, we record several vector field pushforward formulae for the singular coordinate change g(x,y)=xeiyg(x,y)=xe^{iy}. Note gg is simply the polar coordinate transformation (with negative radii permitted). It is a local diffeomorphism away from Z={0}×Z=\{0\}\times\mathbb{R}. We treat a more general type of coordinate change in Section 7, providing additional context for the methods of this section and rendering them somewhat less ad hoc.

Lemma 3.2.

The smooth surjection g:2g:\mathbb{R}^{2}\to\mathbb{C} defined by g(x,y)=xeiyg(x,y)=xe^{iy} satisfies:

g(xx)\displaystyle g_{*}(x\partial_{x}) =xx+yy\displaystyle=x\partial_{x}+y\partial_{y}
g(y)\displaystyle g_{*}(\partial_{y}) =yx+xy\displaystyle=-y\partial_{x}+x\partial_{y}
g(z¯b)\displaystyle g_{*}({{}^{b}\partial}_{\overline{z}}) =z¯z¯\displaystyle=\overline{z}\partial_{\overline{z}}
g(zb)\displaystyle g_{*}({{}^{b}\partial}_{z}) =zz.\displaystyle=z\partial_{z}.

Here, z¯b:-12(xx+iy){{}^{b}\partial}_{\overline{z}}\coloneq\frac{1}{2}(x\partial_{x}+i\partial_{y}), zb:-12(xxiy){{}^{b}\partial}_{z}\coloneq\frac{1}{2}(x\partial_{x}-i\partial_{y}), z¯:-12(x+iy)\partial_{\overline{z}}\coloneq\frac{1}{2}(\partial_{x}+i\partial_{y}), z:-12(xiy)\partial_{z}\coloneq\frac{1}{2}(\partial_{x}-i\partial_{y}). ∎

Secondly, we need the following fact concerning the expression in polar coordinates of plane functions vanishing to infinite order at the origin.

Lemma 3.3.

Define g:2g:\mathbb{R}^{2}\to\mathbb{C} by g(x,y)=xeiyg(x,y)=xe^{iy} and σ:22\sigma:\mathbb{R}^{2}\to\mathbb{R}^{2} by σ(x,y)=(x,y+π)\sigma(x,y)=(-x,y+\pi). Let γ:2\gamma:\mathbb{R}^{2}\to\mathbb{C} be a σ\sigma-invariant (i.e. γσ=γ\gamma\circ\sigma=\gamma) smooth function that vanishes to infinite order on Z:-{0}×Z\coloneq\{0\}\times\mathbb{R}. Then, the pushforward g(γ)g_{*}(\gamma) is a well-defined smooth function on \mathbb{C} that vanishes to infinite order at 0. ∎

Thirdly, we need the following divisibility property of plane functions vanishing to infinite order at the origin.

Lemma 3.4.

Let γ\gamma be a smooth, complex-valued function on \mathbb{C} vanishing to infinite order at 0. Then, (1/z)γ(1/z)\gamma and (1/z¯)γ(1/\overline{z})\gamma extend to smooth, complex-valued functions on \mathbb{C} vanishing to infinite order at 0.∎

We now prove the main result of this section.

Proof of Theorem 3.1.

Since we are only looking for a local solution, there is no harm in assuming γ\gamma satisfies the periodicity assumption γσ=γ\gamma\circ\sigma=\gamma of Lemma 3.3. Indeed, we may take γ\gamma to be compactly-supported in (π2,π2)×(-\frac{\pi}{2},\frac{\pi}{2})\times\mathbb{R}, which is a fundamental domain for σ\sigma, and extend it periodically. This assumption makes the pushforward of LL by g(x,y)=xeiyg(x,y)=xe^{iy} well-defined, as the following computation shows.

g(L)\displaystyle g_{*}(L) =g(z¯b+γzb)\displaystyle=g_{*}({{}^{b}\partial}_{\overline{z}}+\gamma\cdot{{}^{b}\partial}_{z})
=z¯z¯+g(γ)zz\displaystyle=\overline{z}\partial_{\overline{z}}+g_{*}(\gamma)z\partial_{z} (Lemmas 3.2 and 3.3).

Because gg is a local diffeomorphism 2Z{0}\mathbb{R}^{2}\setminus Z\to\mathbb{C}\setminus\{0\}, we have that g(L)g_{*}(L) defines a complex structure on {0}\mathbb{C}\setminus\{0\}. Furthermore, using Lemma 3.4, we can write g(γ)=z¯γg_{*}(\gamma)=\overline{z}\gamma^{\prime} where γ\gamma^{\prime} is another smooth, complex-valued function on \mathbb{C} vanishing to infinite order at 0. Thus,

g(L)=z¯(z¯+γzz).g_{*}(L)=\overline{z}(\partial_{\overline{z}}+\gamma^{\prime}z\partial_{z}).

But now, note that z¯+γzz\partial_{\overline{z}}+\gamma^{\prime}z\partial_{z} defines an (ordinary) complex structure on all of \mathbb{C} (because the γzz\gamma^{\prime}z\partial_{z} term vanishes at 0). So, by the ordinary Newlander-Nirenberg theorem for dimension two, there exists a CC^{\infty}, complex-valued function hh defined near 00\in\mathbb{C} such that

(z¯+γzz)h=0(\partial_{\overline{z}}+\gamma^{\prime}z\partial_{z})h=0

and also satisfying h(0)=0h(0)=0 and xh(0)=1\partial_{x}h(0)=1. Putting f=hgf=h\circ g completes the proof. ∎

4 Proof of main result for dimension 22

This section is devoted to the proof of Theorem 1.2 in the two-dimensional case (n=0n=0, in the notation of Theorem 1.2). Thanks to Mendoza’s Theorem 1.1, we may begin on a complex bb-manifold M=2M=\mathbb{R}^{2} with singular locus Z=×{0}Z=\mathbb{R}\times\{0\} and T0,1bM{{}^{b}T^{0,1}}M spanned by

L:-z¯b+γzb\displaystyle L\coloneq{{}^{b}\partial}_{\overline{z}}+\gamma\cdot{{}^{b}\partial}_{z}

where γ:2\gamma:\mathbb{R}^{2}\to\mathbb{C} is a smooth function vanishing to infinite order on ZZ. Our task is to find new coordinates near (0,0)(0,0) in which T0,1bM{{}^{b}T^{0,1}}M is spanned by z¯b{{}^{b}\partial}_{\overline{z}}.

From Theorem 3.1, there is an open set U2U\subseteq\mathbb{R}^{2} containing (0,0)(0,0) and a smooth function f:Uf:U\to\mathbb{C} such that:

  1. (i)

    Lf=0Lf=0 (bb-holomorphicity),

  2. (ii)

    ff vanishes on ZZ,

  3. (iii)

    fx(0,0)=1f_{x}(0,0)=1 (denoting partial derivatives by subscripts for brevity).

We split ff into its real and imaginary parts

f=u+iv=(u,v).f=u+iv=(u,v).

(recall =2\mathbb{C}=\mathbb{R}^{2} in this article). The key will be the local coordinate change FF defined, roughly speaking, by (u,vu)(u,\frac{v}{u}). In more precise terms, from (ii), we may write

u(x,y)=xa(x,y)\displaystyle u(x,y)=x\cdot a(x,y) v(x,y)=xb(x,y)\displaystyle v(x,y)=x\cdot b(x,y)

where aa and bb are smooth, real-valued functions on UU. From (iii), we have

a(0,0)=ux(0,0)=1\displaystyle a(0,0)=u_{x}(0,0)=1 b(0,0)=vx(0,0)=0.\displaystyle b(0,0)=v_{x}(0,0)=0.

Shrinking UU, we may assume aa is nowhere-vanishing on UU and thus define

F:U2\displaystyle F:U\to\mathbb{R}^{2} F(x,y)=(u(x,y),b(x,y)a(x,y)).\displaystyle F(x,y)=\left(u(x,y),\frac{b(x,y)}{a(x,y)}\right).
Claim 4.1.

After possibly further shrinking UU around (0,0)(0,0), one has:

  1. (a)

    FF is a diffeomorphism from UU onto an open set F(U)2F(U)\subseteq\mathbb{R}^{2},

  2. (b)

    F(0,0)=(0,0)F(0,0)=(0,0) and F(UZ)=F(U)ZF(U\cap Z)=F(U)\cap Z,

  3. (c)

    f=κFf=\kappa\circ F, where κ:22\kappa:\mathbb{R}^{2}\to\mathbb{R}^{2} is given by κ(x,y)=(x,xy)\kappa(x,y)=(x,xy).

Proof.

Statement (c) is a direct consequence of the definition of FF. For statement (b), since uu vanishes on ZZ and ux(0,0)=1u_{x}(0,0)=1, we may, after shrinking UU, assume u1(0)UZu^{-1}(0)\cap U\subseteq Z. For statement (a), it suffices to show the Jacobian of FF at (0,0)(0,0) is invertible. Then, by the inverse function theorem, after possibly shrinking UU again, FF gives a diffeomorphism UF(U)U\to F(U). We claim that the Jacobian of FF at (0,0)(0,0) has the form [101]\begin{bmatrix}1&0\\ *&1\end{bmatrix}. Obviously the top row is correct, so we need only confirm that y(ba)(0,0)=1\partial_{y}\Big{(}\frac{b}{a}\Big{)}(0,0)=1. Away from ZZ, we have

y(ba)=y(vu)=vyuvuyu2.\displaystyle\partial_{y}\Big{(}\frac{b}{a}\Big{)}=\partial_{y}\Big{(}\frac{v}{u}\Big{)}=\frac{v_{y}u-vu_{y}}{u^{2}}.

Collecting real and imaginary parts in Lf=0Lf=0 we have

xuxvy0\displaystyle xu_{x}-v_{y}\sim 0 xvx+uy0,\displaystyle xv_{x}+u_{y}\sim 0,

where \sim denotes agreement of Taylor series at (0,0)(0,0). Thus,

vyuvuy(xux)u+v(xvx)=x2(uxa+vxb)\displaystyle v_{y}u-vu_{y}\sim(xu_{x})u+v(xv_{x})=x^{2}(u_{x}a+v_{x}b)

from which it follows that

y(ba)uxa+vxba2.\displaystyle\partial_{y}\Big{(}\frac{b}{a}\Big{)}\sim\frac{u_{x}a+v_{x}b}{a^{2}}.

In particular, y(ba)(0,0)=(1)(1)+(0)(0)(1)2=1\partial_{y}\Big{(}\frac{b}{a}\Big{)}(0,0)=\frac{(1)(1)+(0)(0)}{(1)^{2}}=1, as was claimed. ∎

It will be relevant to see what the above constructions amount to in the model case when γ=0\gamma=0 and LL is simply z¯b=12(xx+iy){{}^{b}\partial}_{\overline{z}}=\frac{1}{2}(x\partial_{x}+i\partial_{y}). Then, in place of ff, we may use our function

g(x,y)=xeiy=(xcosy,xsiny).g(x,y)=xe^{iy}=(x\cos y,x\sin y).

With this replacement, the diffeomorphism FF above becomes the diffeomorphism GG appearing in Example 2.12

G:×(π2,π2)2\displaystyle G:\mathbb{R}\times(-\tfrac{\pi}{2},\tfrac{\pi}{2})\to\mathbb{R}^{2} G(x,y)=(xcosy,tany).\displaystyle G(x,y)=(x\cos y,\tan y).

Define W:-F(U)W\coloneq F(U) and V:-G1(W)V\coloneq G^{-1}(W), so that we have bb-manifold isomorphisms

(U,UZ){(U,U\cap Z)}(W,WZ){(W,W\cap Z)}(V,VZ).{(V,V\cap Z).}F\scriptstyle{F}G\scriptstyle{G}

Applying the induced maps Fb,Gb{{}^{b}F_{*}},{{}^{b}G_{*}} on bb-tangent bundles (see (2.1) in Section 2), we arrive at two complex bb-structures on (W,WZ)(W,W\cap Z) which can be compared.

Claim 4.2.

The complex bb-structures on WW spanned Fb(L){{}^{b}F_{*}}(L) and Gb(z¯b){{}^{b}G_{*}}({{}^{b}\partial}_{\overline{z}}) are equal.

Proof.

Referring to Proposition 2.7, it suffices to check this away from the singular loci. Thus, we need only check that the usual pushforwards F(L)F_{*}(L) and G(z¯b)G_{*}({{}^{b}\partial}_{\overline{z}}) coincide up to a smooth rescaling over WZW\setminus Z. Note κ(x,y)=(x,xy)\kappa(x,y)=(x,xy) restricts to a diffeomorphism of 2Z\mathbb{R}^{2}\setminus Z and recall f=κFf=\kappa\circ F on UU, g=κGg=\kappa\circ G on VV. Thus, we have diffeomorphisms

UZ{U\setminus Z}κ(WZ){\kappa(W\setminus Z)}VZ,{V\setminus Z,}f\scriptstyle{f}g\scriptstyle{g}

and it suffices to check that (f|UZ)(L)(f|_{U\setminus Z})_{*}(L) and (g|VZ)(z¯b)(g|_{V\setminus Z})_{*}({{}^{b}\partial}_{\overline{z}}) agree up to a smooth rescaling. Indeed, since ff and gg are holomorphic in the usual sense away from ZZ, Proposition 2.13 shows that (f|UZ)(L)(f|_{U\setminus Z})_{*}(L) and (g|VZ)(z¯b)(g|_{V\setminus Z})_{*}({{}^{b}\partial}_{\overline{z}}) are rescalings of z¯=12(x+iy)\partial_{\overline{z}}=\frac{1}{2}(\partial_{x}+i\partial_{y}) (actually, from Lemma 3.2, we even know g(z¯b)=z¯z¯g_{*}({{}^{b}\partial}_{\overline{z}})=\overline{z}\partial_{\overline{z}}). ∎

The preceding claim completes this section’s proof of Theorem 1.2 in the two-dimensional case. The desired local coordinates are provided by G1F:UVG^{-1}\circ F:U\to V.

5 Existence of bb-holomorphic functions: general case

In this section, we extend the results of Section 3 to the general case. That is, we show that nontrivial bb-holomorphic functions exist near points on the singular locus of any complex bb-manifold. By Theorem 1.1, this amounts to the following.

Theorem 5.1.

Consider a complex bb-manifold M=n+1=2n+2M=\mathbb{C}^{n+1}=\mathbb{R}^{2n+2} with singular locus Z={0}×2n+1Z=\{0\}\times\mathbb{R}^{2n+1} equipped with coordinates (x0,y0,,xn,yn)=(z0,,zn)(x_{0},y_{0},\ldots,x_{n},y_{n})=(z_{0},\ldots,z_{n}) with complex bb-structure T0,1bM{{}^{b}T^{0,1}}M spanned by

L0\displaystyle L_{0} :-z¯b+Γ0\displaystyle\coloneq{{}^{b}\partial}_{\overline{z}}+\Gamma_{0}
Lj\displaystyle L_{j} :-z¯j+Γj\displaystyle\coloneq\partial_{\overline{z}_{j}}+\Gamma_{j} j=1,,n\displaystyle j=1,\ldots,n

where Γ0,,Γn\Gamma_{0},\ldots,\Gamma_{n} are linear combinations of z0b,z1,,zn{{}^{b}\partial}_{z_{0}},\partial_{z_{1}},\ldots,\partial_{z_{n}} with coefficients in smooth, complex-valued functions on MM vanishing to infinite order on ZZ. Then, there exists an open set UMU\subseteq M containing 0 and bb-holomorphic functions f0,f1,,fn:Uf_{0},f_{1},\ldots,f_{n}:U\to\mathbb{C} such that:

  1. (i)

    f0f_{0} vanishes on ZZ and fj(0)=0f_{j}(0)=0 for j=1,,nj=1,\ldots,n,

  2. (ii)

    xjfk(0)=δj,k\partial_{x_{j}}f_{k}(0)=\delta_{j,k} for j,k{0,1,,n}j,k\in\{0,1,\ldots,n\}, using Kronecker delta notation.

The proof is largely the same as Theorem 3.1, so we will be somewhat brief. First, we state straightforward higher-dimensional generalizations of Lemmas 3.3 and 3.4. Again, we omit the proofs of these statements.

Lemma 5.2.

Define

g\displaystyle g :2\displaystyle:\mathbb{R}^{2}\to\mathbb{C} g(x,y)\displaystyle g(x,y) =xeiy\displaystyle=xe^{iy} g~\displaystyle\widetilde{g} :2n+2n+1\displaystyle:\mathbb{R}^{2n+2}\to\mathbb{C}^{n+1} g~=g×id\displaystyle\widetilde{g}=g\times\mathrm{id}
σ\displaystyle\sigma :22\displaystyle:\mathbb{R}^{2}\to\mathbb{R}^{2} σ(x,y)\displaystyle\sigma(x,y) =(x,y+π)\displaystyle=(-x,y+\pi) σ~\displaystyle\widetilde{\sigma} :2n+22n+2\displaystyle:\mathbb{R}^{2n+2}\to\mathbb{R}^{2n+2} σ~=σ×id\displaystyle\widetilde{\sigma}=\sigma\times\mathrm{id}

Let γ:2n+2\gamma:\mathbb{R}^{2n+2}\to\mathbb{C} be a σ~\widetilde{\sigma}-invariant smooth function that vanishes to infinite order on Z:-{0}×2n+1Z\coloneq\{0\}\times\mathbb{R}^{2n+1}. Then, the pushforward g~(γ)\widetilde{g}_{*}(\gamma) is a well-defined smooth function on n+1\mathbb{C}^{n+1} that vanishes to infinite order on Z:-{0}×nZ^{\prime}\coloneq\{0\}\times\mathbb{C}^{n}. ∎

Corollary 5.3.

Suppose γ0,,γn:2n+2\gamma_{0},\ldots,\gamma_{n}:\mathbb{R}^{2n+2}\to\mathbb{C} are σ~\widetilde{\sigma}-invariant smooth functions vanishing infinite order on Z={0}×2n+1Z=\{0\}\times\mathbb{R}^{2n+1}. Define Γ=γ0z¯0b+j=1nγjz¯j\Gamma=\gamma_{0}\cdot{{}^{b}\partial}_{\overline{z}_{0}}+\sum_{j=1}^{n}\gamma_{j}\cdot\partial_{\overline{z}_{j}} (note Γ\Gamma is σ~\widetilde{\sigma}-invariant in the sense that σ~(Γ)=Γ\widetilde{\sigma}_{*}(\Gamma)=\Gamma). Then, the pushforward g~(Γ)\widetilde{g}_{*}(\Gamma) is a well-defined smooth vector field on n+1\mathbb{C}^{n+1} vanishing to infinite order on Z={0}×nZ^{\prime}=\{0\}\times\mathbb{C}^{n}.

Proof.

From Lemma 3.2 and Lemma 5.2, we have g~(Γ)=g~(γ0)z¯0z¯0+j=1ng~(γj)z¯j\widetilde{g}_{*}(\Gamma)=\widetilde{g}_{*}(\gamma_{0})\overline{z}_{0}\partial_{\overline{z}_{0}}+\sum_{j=1}^{n}\widetilde{g}_{*}(\gamma_{j})\partial_{\overline{z}_{j}}. ∎

Lemma 5.4.

Let n+1\mathbb{C}^{n+1} have coordinates (z0,,zn)(z_{0},\ldots,z_{n}). Suppose γ\gamma is a smooth, complex-valued function on n+1\mathbb{C}^{n+1} vanishing to infinite order on Z:-{0}×nZ^{\prime}\coloneq\{0\}\times\mathbb{C}^{n}. Then, (1/z0)γ(1/z_{0})\gamma and (1/z¯0)γ(1/\overline{z}_{0})\gamma extend to smooth, complex-valued functions on n+1\mathbb{C}^{n+1} vanishing to infinite order on ZZ^{\prime}. ∎

Having made the above preparations, we now proceed to the proof of this section’s main result.

Proof of Theorem 5.1.

As in the proof of Lemma 3.1, since we only desire local bb-holomorphic functions, there is no harm assuming the deformations terms Γj\Gamma_{j} to be σ~\widetilde{\sigma}-invariant. With this assumption, Lemma 3.2 and Corollary 5.3 together imply that the vector fields LjL_{j} and Γj\Gamma_{j} have well-defined pushforwards by g~\widetilde{g}. Indeed,

g~(L0)\displaystyle\widetilde{g}_{*}(L_{0}) =z¯0z¯0+g~(Γ0)\displaystyle=\overline{z}_{0}\partial_{\overline{z}_{0}}+\widetilde{g}_{*}(\Gamma_{0})
g~(Lj)\displaystyle\widetilde{g}_{*}(L_{j}) =z¯j+g~(Γj)\displaystyle=\partial_{\overline{z}_{j}}+\widetilde{g}_{*}(\Gamma_{j}) j=1,,n,\displaystyle j=1,\ldots,n,

where g~(Γj)\widetilde{g}_{*}(\Gamma_{j}) is a smooth vector field on n+1\mathbb{C}^{n+1} vanishing to infinite order on Z:-{0}×nZ^{\prime}\coloneq\{0\}\times\mathbb{C}^{n} for j=0,,nj=0,\ldots,n.

Because g~\widetilde{g} is a local diffeomorphism 2n+2Zn+1Z\mathbb{R}^{2n+2}\setminus Z\to\mathbb{C}^{n+1}\setminus Z^{\prime} and bracket commutes with pushforward, g~(Lj)\widetilde{g}_{*}(L_{j}), j=0,,nj=0,\ldots,n at least define a complex structure on n+1Z\mathbb{C}^{n+1}\setminus Z^{\prime}. Going further, we can use Lemma 5.4 to write g~(Γ0)=z¯0Γ0\widetilde{g}_{*}(\Gamma_{0})=\overline{z}_{0}\Gamma_{0}^{\prime}, where Γ0\Gamma_{0}^{\prime} is another smooth vector field on n+1\mathbb{C}^{n+1} vanishing to infinite order on ZZ^{\prime}. We claim that

1z¯0g~(L0)\displaystyle\tfrac{1}{\overline{z}_{0}}\widetilde{g}_{*}(L_{0}) =z¯0+Γ0\displaystyle=\partial_{\overline{z}_{0}}+\Gamma_{0}^{\prime} (5.1)
g~(Lj)\displaystyle\widetilde{g}_{*}(L_{j}) =z¯j+g~(Γj)\displaystyle=\partial_{\overline{z}_{j}}+\widetilde{g}_{*}(\Gamma_{j}) j=1,,n\displaystyle j=1,\ldots,n (5.2)

define an ordinary complex structure on the whole of n+1\mathbb{C}^{n+1} (extending the aforementioned complex structure on n+1Z\mathbb{C}^{n+1}\setminus Z^{\prime}). Indeed, the deformation terms vanish on ZZ^{\prime}, so the vector fields (5.1) together with their complex conjugates, form a global frame for the complexified tangent bundle. By a simple computation using the derivation property of the Lie bracket, involutivity holds away from ZZ^{\prime}, hence everywhere by a continuity argument.

Now, by an application of the classical Newlander-Nirenberg theorem, there is an open neighbourhood Un+1U\subseteq\mathbb{C}^{n+1} of 0 and smooth functions h0,,hn:Uh_{0},\ldots,h_{n}:U\to\mathbb{C} that are holomorphic for the complex structure defined by (5.1) and furthermore satisfy hj(0)=0h_{j}(0)=0, xjhk(0)=δj,k\partial_{x_{j}}h_{k}(0)=\delta_{j,k} for j,k{0,1,,n}j,k\in\{0,1,\ldots,n\}. In fact, again using the vanishing of the deformation terms along ZZ^{\prime}, the complex structure on n+1\mathbb{C}^{n+1} defined by (5.1) is such that Z={0}×nZ^{\prime}=\{0\}\times\mathbb{C}^{n} sits as a complex submanifold. Indeed, the inherited complex structure on ZZ^{\prime} is its standard one spanned by z¯1,,z¯n\partial_{\overline{z}_{1}},\ldots,\partial_{\overline{z}_{n}}. Using the implicit function theorem for several complex variables, complex submanifolds of complex codimension-1 can locally be written as preimages of regular values of holomorphic functions. Thus, possibly shrinking UU, we may even choose h0h_{0} to vanish on ZZ^{\prime}. Putting fj=hjg~f_{j}=h_{j}\circ\widetilde{g} for j=0,1,,nj=0,1,\ldots,n completes the proof of the result. ∎

6 The bb-Newlander-Nirenberg theorem: general case

This section is devoted to the proof of our main result Theorem 1.2. In light of Mendoza’s Theorem 1.1, we may begin on a complex bb-manifold M=n+1=2n+2M=\mathbb{C}^{n+1}=\mathbb{R}^{2n+2} with singular locus Z={0}×2n+1Z=\{0\}\times\mathbb{R}^{2n+1} equipped with coordinates (x0,y0,,xn,yn)=(z0,,zn)(x_{0},y_{0},\ldots,x_{n},y_{n})=(z_{0},\ldots,z_{n}) with complex bb-structure T0,1bM{{}^{b}T^{0,1}}M spanned by

L0\displaystyle L_{0} :-z¯b+Γ0\displaystyle\coloneq{{}^{b}\partial}_{\overline{z}}+\Gamma_{0}
Lj\displaystyle L_{j} :-z¯j+Γj\displaystyle\coloneq\partial_{\overline{z}_{j}}+\Gamma_{j} j=1,,n\displaystyle j=1,\ldots,n

where Γ0,,Γn\Gamma_{0},\ldots,\Gamma_{n} are smooth, complex vector fields on MM vanishing to infinite order on ZZ.

By Theorem 5.1, there is an open set U2n+2U\subseteq\mathbb{R}^{2n+2} containing 0 and smooth functions f0,,fn:Uf_{0},\ldots,f_{n}:U\to\mathbb{C} satisfying:

  1. (i)

    Ljfk=0L_{j}f_{k}=0 for j,k{0,1,,n}j,k\in\{0,1,\ldots,n\} (bb-holomorphicity),

  2. (ii)

    f0f_{0} vanishes on ZZ and fj(0)=0f_{j}(0)=0 for j=1,,nj=1,\ldots,n,

  3. (iii)

    xjfk(0)=δj,k\partial_{x_{j}}f_{k}(0)=\delta_{j,k} for j,k{0,1,,n}j,k\in\{0,1,\ldots,n\}.

As in Section 4, we decompose f0f_{0} into real and imaginary parts f0=u+ivf_{0}=u+iv. Since uu and vv vanish on ZZ, we may write u=x0au=x_{0}\cdot a and v=x0bv=x_{0}\cdot b, where aa and bb are real-valued smooth functions. As in Section 4, by shrinking UU, we may enforce that aa is nowhere vanishing on UU and that:

  1. (a)

    F=(a,ba,f1,,fn)F=(a,\tfrac{b}{a},f_{1},\ldots,f_{n}) is a diffeomorphism from UU onto an open set F(U)2n+2F(U)\subseteq\mathbb{R}^{2n+2},

  2. (b)

    F(0)=0F(0)=0 and F(UZ)=F(U)ZF(U\cap Z)=F(U)\cap Z,

  3. (c)

    f=κ~Ff=\widetilde{\kappa}\circ F where f=(f0,f1,,fn)f=(f_{0},f_{1},\ldots,f_{n}) and κ~=κ×id:2n+22n+2\widetilde{\kappa}=\kappa\times\mathrm{id}:\mathbb{R}^{2n+2}\to\mathbb{R}^{2n+2} where κ:22\kappa:\mathbb{R}^{2}\to\mathbb{R}^{2} is κ(x,y)=(x,xy)\kappa(x,y)=(x,xy).

As in Section 4, we argue that the pushforward by FF of the complex bb-structure on UU spanned by L0,,LnL_{0},\ldots,L_{n} is independent of the deformation terms Γ0,,Γn\Gamma_{0},\ldots,\Gamma_{n}. Again, by Proposition 2.7, if suffices check this away from the singular locus ZZ. Using the fact that κ~\widetilde{\kappa} restricts to a self-diffeomorphism of 2n+2Z\mathbb{R}^{2n+2}\setminus Z, we have that ff defines a diffeomorphism of UZU\setminus Z onto κ(F(UZ))\kappa(F(U\setminus Z)). Because the components of ff are holomorphic for the induced complex structure on UZU\setminus Z, Proposition 2.13 implies that f|UZf|_{U\setminus Z} pushes forward the subbundle spanned by L0,,LnL_{0},\ldots,L_{n} to the subbundle spanned by z¯0,,z¯n\partial_{\overline{z}_{0}},\ldots,\partial_{\overline{z}_{n}}. Thus, the pushforward by F|UZF|_{U\setminus Z} of the subbundle spanned by L0,,LnL_{0},\ldots,L_{n} is completely determined to be the pullback of the standard complex structure on κ(F(UZ))\kappa(F(U\setminus Z)) by κ\kappa. Running the same argument in the case where the deformation terms are all zero, we obtain the desired coordinate change.

7 Remarks on singular pushforwards

The polar coordinate change g:2g:\mathbb{R}^{2}\to\mathbb{C}, g(x,y)=xeiyg(x,y)=xe^{iy} played a crucial role in this article by enabling us to relate the nonelliptic operator zb=12(xxiy){{}^{b}\partial}_{z}=\frac{1}{2}(x\partial_{x}-i\partial_{y}) to the standard complex partial derivative operator z=12(xiy)\partial_{z}=\frac{1}{2}(\partial_{x}-i\partial_{y}) by way of the elementary, but perhaps somewhat mysterious, pushforward formula g(zb)=zzg_{*}({{}^{b}\partial}_{z})=z\partial_{z}. In this section, we shed additional light on this pushforward formula by placing it in a more general context. The material in this section will also play a role in planned future work on the bb-Newlander-Nirenberg theorem for complex bkb^{k}-manifolds [1].

Suppose that, more generally, we wish to relate the vector field

12(xkxiy)\displaystyle\tfrac{1}{2}(x^{k}\partial_{x}-i\partial_{y})

to z\partial_{z}, for kk a positive integer. More generally still, suppose we wish to relate

12(f(x)xiy)\displaystyle\tfrac{1}{2}(f(x)\partial_{x}-i\partial_{y})

to z\partial_{z}, where f(z)f(z) is an entire function on \mathbb{C} whose restriction f(x)f(x) to \mathbb{R} is real-valued.

We first recall a few more-or-less notational points concerning holomorphic vector fields and imaginary-time flows. One may refer to [3], pp. 39 for additional details. Let ff be a holomorphic function on \mathbb{C} so that V:-f(z)zV\coloneq f(z)\partial_{z} is a holomorphic vector field on \mathbb{C}. We may decompose VV as

V=12(XiJX),V=\tfrac{1}{2}(X-iJX),

where the real vector fields X=Re(f)x+Im(f)yX=\mathrm{Re}(f)\partial_{x}+\mathrm{Im}(f)\partial_{y} and JX=Im(f)x+Re(f)yJX=-\mathrm{Im}(f)\partial_{x}+\mathrm{Re}(f)\partial_{y} commute with one another. We may then define the complex-time flow of VV by

ϕwV(z):-ϕsXϕtJX(z)\displaystyle\phi^{V}_{w}(z)\coloneq\phi^{X}_{s}\circ\phi^{JX}_{t}(z) where w=s+it,\displaystyle\text{where }w=s+it,

with the usual caveat that the flow need only be defined for ww sufficiently close to 0.

This flow is jointly holomorphic in ww and zz. One practical consequence of this joint holomorphicity is that, if f(z)f(z) is real on the xx-axis, so that XX coincides with f(x)xf(x)\partial_{x} on the xx-axis, then the complex-time flow of the holomorphic vector field f(z)zf(z)\partial_{z} may be obtained by analytic continuation in both variables of the real-time flow of the one-dimensional vector field f(x)xf(x)\partial_{x}.

Example 7.1.

The real-time flow of the one-dimensional vector field X1:-xxX_{1}\coloneq x\partial_{x} is given by ϕtX1(x)=etx\phi^{X_{1}}_{t}(x)=e^{t}x. Accordingly, the complex-time flow of the holomorphic vector field V1:-zzV_{1}\coloneq z\partial_{z} is given by ϕwV1(z)=ewz\phi^{V_{1}}_{w}(z)=e^{w}z.

Example 7.2.

The real-time flow of the one-dimensional vector field X2:-x2xX_{2}\coloneq x^{2}\partial_{x} is given by ϕtX2(x)=x1tx\phi^{X_{2}}_{t}(x)=\frac{x}{1-tx}. Accordingly, the complex-time flow of the holomorphic vector field V2:-z2zV_{2}\coloneq z^{2}\partial_{z} is given by ϕwV2(z)=z1wz\phi^{V_{2}}_{w}(z)=\frac{z}{1-wz}.

Example 7.3.

For any integer k2k\geq 2, the real-time flow of the one-dimensional vector field Xk:-xkxX_{k}\coloneq x^{k}\partial_{x} is given by ϕtXk(x)=x1(k1)txk1k1\phi^{X_{k}}_{t}(x)=\frac{x}{\sqrt[k-1]{1-(k-1)tx^{k-1}}}. Accordingly, the complex-time flow of the holomorphic vector field Vk:-zkzV_{k}\coloneq z^{k}\partial_{z} is given by ϕwVk(z)=z1(k1)wzk1k1\phi^{V_{k}}_{w}(z)=\frac{z}{\sqrt[k-1]{1-(k-1)wz^{k-1}}}.

Bearing in mind the above notations, we now give the main result of this section.

Proposition 7.4.

Let ff be a holomorphic function on \mathbb{C} that is real-valued on \mathbb{R}. Let U2U\subseteq\mathbb{R}^{2} be a sufficiently small open neighbourhood of ×{0}\mathbb{R}\times\{0\} that h:Uh:U\to\mathbb{C}, h(x,y)=ϕiyf(z)z(x)h(x,y)=\phi^{f(z)\partial_{z}}_{iy}(x) is defined. Then, the vector field 12(f(x)x+iy)\frac{1}{2}(f(x)\partial_{x}+i\partial_{y}) on UU is hh-related to the holomorphic vector field f(z)zf(z)\partial_{z} on \mathbb{C}.

Proof.

Decompose V=f(z)zV=f(z)\partial_{z} in the form V=12(XiJX)V=\frac{1}{2}(X-iJX), as above. Thus, by definition, h(x,y)=ϕyJX(x)h(x,y)=\phi^{JX}_{y}(x). For fixed (x,y)U(x,y)\in U, taking tt sufficiently small, we have

h(x,y+t)=ϕy+tJX(x)=ϕtJX(h(x,y+t)).h(x,y+t)=\phi^{JX}_{y+t}(x)=\phi^{JX}_{t}(h(x,y+t)).

Thus, the vector field y\partial_{y} is hh-related to JXJX. Similarly,

ϕtXh(x,y)=ϕtXϕyJX(x)=ϕyJXϕtX(x)=ϕyJXϕtf(x)x(x)=h(ϕtf(x)x(x),y)=h(ϕtf(x)x(x,y))\phi^{X}_{t}h(x,y)=\phi^{X}_{t}\phi^{JX}_{y}(x)=\phi^{JX}_{y}\phi^{X}_{t}(x)=\phi^{JX}_{y}\phi^{f(x)\partial_{x}}_{t}(x)=h(\phi^{f(x)\partial_{x}}_{t}(x),y)=h(\phi^{f(x)\partial_{x}}_{t}(x,y))

Thus, the vector field f(x)xf(x)\partial_{x} is hh-related to XX. The result follows by linearity. ∎

Corollary 7.5.

Let g2:2g_{2}:\mathbb{R}^{2}\to\mathbb{C}, g2(x,y)=x1ixyg_{2}(x,y)=\frac{x}{1-ixy}. Then, 12(x2x+iy)\frac{1}{2}(x^{2}\partial_{x}+i\partial_{y}) is g2g_{2}-related to z2zz^{2}\partial_{z}.

Corollary 7.6.

Fix an integer k2k\geq 2. Let gk:2g_{k}:\mathbb{R}^{2}\to\mathbb{C}, gk(x,y)=x1i(k1)xk1yk1g_{k}(x,y)=\frac{x}{\sqrt[k-1]{1-i(k-1)x^{k-1}y}}. Then 12(xkxiy)\tfrac{1}{2}(x^{k}\partial_{x}-i\partial_{y}) is gkg_{k}-related to zkzz^{k}\partial_{z}.

We remark that gk(x,y)=x1i(k1)xk1yk1g_{k}(x,y)=\frac{x}{\sqrt[k-1]{1-i(k-1)x^{k-1}y}} collapses Z={0}×Z=\{0\}\times\mathbb{R} to {0}\{0\} and defines a diffeomorphism 2Z{reiθ:r0,π2(k1)<θ<π2(k1)}\mathbb{R}^{2}\setminus Z\to\{re^{i\theta}\in\mathbb{C}:r\neq 0,-\frac{\pi}{2(k-1)}<\theta<\frac{\pi}{2(k-1)}\}. Note that the line Re(z)=1\mathrm{Re}(z)=1 is completely contained in the principal branch of zk1\sqrt[k-1]{z}. For purposes of clarification, sample plots of the real vector field JXJX in the decomposition zkz=12(XiJX)z^{k}\partial_{z}=\frac{1}{2}(X-iJX) are given below. Note all integral curves passing through the xx-axis are complete.

Refer to caption
Refer to caption
Refer to caption
Figure 1: The vector field JX=Im(zk)x+iRe(zk)yJX=-\mathrm{Im}(z^{k})\partial_{x}+i\mathrm{Re}(z^{k})\partial_{y} for k=2,3,4k=2,3,4.

References

  • [1] T. Barron and M. Francis. On automorphisms of complex bkb^{k}-manifolds, 2023 (arXiv preprint).
  • [2] M. Braverman, Y. Loizides, and Y. Song. Geometric quantization of bb-symplectic manifolds. J. Symplectic Geom., 19(1):1–36, 2021.
  • [3] F. Forstnerič. Stein manifolds and holomorphic mappings, volume 56 of Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics]. Springer, Cham, second edition, 2017. The homotopy principle in complex analysis.
  • [4] V. Guillemin, E. Miranda, and A. R. Pires. Symplectic and Poisson geometry on bb-manifolds. Adv. Math., 264:864–896, 2014.
  • [5] Y. Lin, Y. Loizides, R. Sjamaar, and Y. Song. Symplectic reduction and a darboux-moser-weinstein theorem for lie algebroids, 2022.
  • [6] R. B. Melrose. The Atiyah-Patodi-Singer index theorem, volume 4 of Research Notes in Mathematics. A K Peters, Ltd., Wellesley, MA, 1993.
  • [7] G. A. Mendoza. Complex bb-manifolds. In Geometric and spectral analysis, volume 630 of Contemp. Math., pages 139–171. Amer. Math. Soc., Providence, RI, 2014.
  • [8] A. Newlander and L. Nirenberg. Complex analytic coordinates in almost complex manifolds. Ann. of Math. (2), 65:391–404, 1957.