The Newlander-Nirenberg theorem for
complex -manifolds
Abstract
Melrose defined the -tangent bundle of a smooth manifold with boundary as the vector bundle whose sections are vector fields on tangent to the boundary. Mendoza defined a complex -manifold as a manifold with boundary together with an involutive splitting of the complexified -tangent bundle into complex conjugate factors. In this article, we prove complex -manifolds have a single local model depending only on dimension. This can be thought of as the Newlander-Nirenberg theorem for complex -manifolds: there are no “local invariants”. Our proof uses Mendoza’s existing result that complex -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 and use coordinates 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 is an involutive subbundle of the complexified tangent bundle such that where . The Newlander-Nirenberg theorem implies that every point belongs to a coordinate chart in which is spanned by
Accordingly, is spanned by
In this article, a -manifold is a smooth manifold equipped with a closed hypersurface referred to as the singular locus. The -tangent bundle is the vector bundle over whose smooth sections are the vector fields on that are tangent along . These terminologies stem from the -geometry of Melrose [6], also called log-geometry.
Mendoza [7] defined a complex -structure on a -manifold to be an involutive subbundle of the complexified -tangent bundle such that , where . Thus, complex -structures are defined exactly analogously to complex structures, replacing the tangent bundle by the -tangent bundle.
To complete the analogy, one would like there to be a Newlander-Nirenberg type result to the effect that all complex -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 -manifold in terms of an appropriately defined -holomorphic atlas. A logical candidate model for a complex -manifolds is , equipped with coordinates with spanned by
Accordingly, is spanned by
Our goal in this article is to show that, locally near points on the singular locus, every complex -manifold of dimension 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 , let be a -dimensional complex -manifold with singular locus . Then, every has an open neighbourhood on which there are smooth local coordinates centered at with vanishing on and smooth, complex vector fields on vanishing to infinite order on such that
is a frame for over . Moreover, each of is a linear combination of whose coefficients are smooth, complex-valued functions on vanishing to infinite order on .
The contribution of the present article will be to improve the above result by showing that the deformation terms can be got rid of. Thus, our main result is the following.
Theorem 1.2.
With , let be a -dimensional complex -manifold with singular locus . Then, every has an open neighbourhood on which there are smooth local coordinates centered at with vanishing on such that
is a frame for over .
Readers familiar with the techniques needed to prove the classical Newlander-Nirenberg theorem will be unsurprised that nonellipticity of the operator 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 , (negative radii permitted). Notice that , i.e. is “-holomorphic”.
Let us summarize the contents of this paper. In Section 2, we collect definitions and basic results on complex -manifolds. In Section 3, we prove in the two-dimensional case that nontrivial -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 in a more general context which will be relevant for planned future work on complex -manifolds (see [1] for the definition of a complex -manifold).
2 Background
The language of -geometry, also known as log-geometry, was introduced by Melrose [6]. One encounters two superficially different formulations of -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 -manifold is a smooth manifold together with a closed hypersurface which we refer to as the singular locus. An isomorphism of -manifolds is a diffeomorphism that preserves the singular loci.
Note ordinary smooth manifolds may be considered as -manifolds, taking .
Definition 2.2.
A -vector field on a -manifold is a smooth vector field on that is tangent to . The collection of -vector fields on is denoted .
Example 2.3.
Consider the -manifold with coordinates and singular locus . Then, , the free -module generated by .
The example above completely captures the local structure of -manifolds. Accordingly, for any -manifold , is a projective -module closed under Lie bracket. Applying Serre-Swan duality to the inclusion , one has a corresponding Lie algebroid whose anchor map, denoted , induces (abusing notation) an identification .
Definition 2.4.
For a -manifold , the Lie algebroid satisfying described above is called the -tangent bundle of .
In Example 2.3, is a global frame for . The anchor map descends from evaluation of vector fields. Thus, over , is an isomorphism and, over , the kernel of is the line bundle spanned by . In general, for any -manifold , the bundles and are canonically isomorphic over via .
If is an isomorphism of -manifolds, it is clear that preserves the associated modules of -vector fields. By Serre-Swan duality, induces a Lie algebroid isomorphism
(2.1) |
Over , if we apply the aforementioned natural identifications of the -tangent bundle and usual tangent bundle, coincides with , the usual induced isomorphism given by pushforward of tangent vectors.
Definition 2.5.
The -cotangent bundle of a -manifold is the dual of the -tangent bundle .
The dual of the anchor map is likewise an isomorphism when restricted over . In particular, is injective on a dense open set, and so (what is equivalent) induces an injective mapping . Put in simpler terms, a 1-form is fully determined by its pairing with -vector fields. The -exterior derivative of a smooth function on is defined as:
(2.2) |
Actually, one may quite legitimately regard and as identical, with the latter notation merely hinting that one only intends to pair the form with -vector fields.
A general philosophy of -calculus is that many classical geometries admit -analogues in which the role of tangent bundle is played by the -tangent bundle. A notable example is -symplectic geometry ([4], [5]). Complex -geometry was introduced by Mendoza [7] who furthermore made a systematic study of the -Dolbeault complex. We remark that the basic idea of a complex -structure is mentioned in passing on pp. 218 of [6].
Definition 2.6.
A complex -structure on an (even-dimensional) -manifold is a complex subbundle of the complexified -tangent bundle satisfying:
-
(i)
, where ,
-
(ii)
is involutive.
A complex -manifold is a -manifold equipped with a complex -structure. An isomorphism of complex -manifolds is an isomorphism of their underlying -manifolds satisfying . Here, by abuse of notation, denotes the complexification of the isomorphism defined above (2.1).
Because is naturally isomorphic to away from , a -complex structure on in particular gives a complex structure in the usual sense on . In particular, one may consider ordinary complex manifolds as special cases of complex -manifolds with .
Actually, for any complex -manifold , the restricted complex structure on determines the -complex structure, as the following proposition shows. Of course, not all complex structures on are the restriction of a (unique) -complex structure on , just those that degenerate in a particular way at .
Proposition 2.7.
Let be a complex -manifold with singular locus for . Suppose that is a diffeomorphism such that:
-
(i)
(i.e. is an isomorphism of -manifolds),
-
(ii)
restricts to an isomorphism of (ordinary) complex manifolds .
Then, is an isomorphism of complex -manifolds.
Proof.
From (i), we have an induced isomorphism on the complexified -tangent bundles . From (ii), and applying the natural isomorphisms of -tangent bundle and ordinary tangent bundle away from the singular loci, the subbundles and of agree away from . The conclusion follows by a continuity argument. ∎
Taking duals, the splitting given by a complex -structure induces an associated splitting of the complexified -cotangent bundle and of the complexified -exterior derivative , as tabulated below:
In effect, , respectively , is restricted to the --vector fields, respectively the --vector fields (that is to say, sections of , respectively sections of ).
Definition 2.8.
A -holomorphic function on a complex -manifold (or an open subset thereof) is a smooth function satisfying . We write for the collection of -holomorphic functions on .
In more concrete terms, is -holomorphic if for every --vector field or (what is sufficient), all in a given frame for . Because is involutive, is a ring with respect to pointwise-multiplication.
Example 2.9.
Consider the -manifold with singular locus . Then, , the free -module generated by and . An example of a complex -structure for has spanned by
and, correspondingly, spanned by
A function is -holomorphic precisely when . An important globally-defined -holomorphic function on was mentioned in the introduction
More generally, if is a usual holomorphic function defined near , then is a -holomorphic function defined near . There also exist -holomorphic functions defined near not of the latter form and, indeed, not real-analytic near . For example, on the domain , one has a -holomorphic defined by for and for . These examples are also noted in [1], Example 8.1.
Remark 2.10.
Observe that the -holomorphic function in the above example may be conceptualized as , where is the flow of . This somewhat cryptic remark will be expanded on in Section 7.
Example 2.11.
Consider the -manifold with as its singular locus. Introduce coordinates . Thus,
An example of a complex -structure for is the one spanned by:
Correspondingly, a frame for is:
The function from Example 2.9 is still -holomorphic, as are the complex coordinate functions , . Additional -holomorphic functions may be obtained by applying an -variable holomorphic function (in the usual sense) to .
Example 2.12.
Consider the complex -manifold with singular locus and with spanned by
This complex -manifold is isomorphic to a neighbourhood of the origin of the complex -manifold with spanned by from Example 2.9. Indeed, the diffeomorphism defined by satisfies . Thus, -holomorphic functions on may be obtained by pushforward through . For example, the pushforward of by is which indeed satisfies .
For a general complex -manifold , it is not immediately clear whether, locally near points on the singular locus , any nonconstant -holomorphic functions must exist at all. This article will confirm that they do (Sections 3 and 5).
Note that a -holomorphic function on a complex -manifold restricts to a holomorphic function in the usual sense on the complex manifold . Indeed, in a similar spirit to Proposition 2.7, if is smooth and restricts to a holomorphic function on , then it is -holomorphic on 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 be a complex manifold. Let and be open. Suppose are holomorphic functions such that , defines a diffeomorphism. Then, maps onto over where, by definition, the latter is spanned by , .
Proof.
Let be a -vector field defined on . Using for and (the th coordinate function) for , we obtain
If we write , where are smooth, complex-valued functions on , then the above gives for . Thus, we obtain
so that is a vector field on . ∎
In the case of complex -manifolds, -holomorphic functions cannot directly play the role of coordinate functions of charts. To see this, note that any solution of (where ) is necessarily constant along .
3 Existence of -holomorphic functions: dimension
In this section, we show that, near any point on the singular locus of a two-dimensional complex -manifold, there is a nontrivial -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 , ).
Theorem 3.1.
Consider a complex -manifold with singular locus and with spanned by
where is a smooth function vanishing to infinite order on . Then, there is a neighbourhood of and a -holomorphic function that vanishes on and satisfies .
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 . Note is simply the polar coordinate transformation (with negative radii permitted). It is a local diffeomorphism away from . 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 defined by satisfies:
Here, , , , . ∎
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 by and by . Let be a -invariant (i.e. ) smooth function that vanishes to infinite order on . Then, the pushforward is a well-defined smooth function on that vanishes to infinite order at . ∎
Thirdly, we need the following divisibility property of plane functions vanishing to infinite order at the origin.
Lemma 3.4.
Let be a smooth, complex-valued function on vanishing to infinite order at . Then, and extend to smooth, complex-valued functions on vanishing to infinite order at .∎
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 satisfies the periodicity assumption of Lemma 3.3. Indeed, we may take to be compactly-supported in , which is a fundamental domain for , and extend it periodically. This assumption makes the pushforward of by well-defined, as the following computation shows.
(Lemmas 3.2 and 3.3). |
Because is a local diffeomorphism , we have that defines a complex structure on . Furthermore, using Lemma 3.4, we can write where is another smooth, complex-valued function on vanishing to infinite order at . Thus,
But now, note that defines an (ordinary) complex structure on all of (because the term vanishes at ). So, by the ordinary Newlander-Nirenberg theorem for dimension two, there exists a , complex-valued function defined near such that
and also satisfying and . Putting completes the proof. ∎
4 Proof of main result for dimension
This section is devoted to the proof of Theorem 1.2 in the two-dimensional case (, in the notation of Theorem 1.2). Thanks to Mendoza’s Theorem 1.1, we may begin on a complex -manifold with singular locus and spanned by
where is a smooth function vanishing to infinite order on . Our task is to find new coordinates near in which is spanned by .
From Theorem 3.1, there is an open set containing and a smooth function such that:
-
(i)
(-holomorphicity),
-
(ii)
vanishes on ,
-
(iii)
(denoting partial derivatives by subscripts for brevity).
We split into its real and imaginary parts
(recall in this article). The key will be the local coordinate change defined, roughly speaking, by . In more precise terms, from (ii), we may write
where and are smooth, real-valued functions on . From (iii), we have
Shrinking , we may assume is nowhere-vanishing on and thus define
Claim 4.1.
After possibly further shrinking around , one has:
-
(a)
is a diffeomorphism from onto an open set ,
-
(b)
and ,
-
(c)
, where is given by .
Proof.
Statement (c) is a direct consequence of the definition of . For statement (b), since vanishes on and , we may, after shrinking , assume . For statement (a), it suffices to show the Jacobian of at is invertible. Then, by the inverse function theorem, after possibly shrinking again, gives a diffeomorphism . We claim that the Jacobian of at has the form . Obviously the top row is correct, so we need only confirm that . Away from , we have
Collecting real and imaginary parts in we have
where denotes agreement of Taylor series at . Thus,
from which it follows that
In particular, , as was claimed. ∎
It will be relevant to see what the above constructions amount to in the model case when and is simply . Then, in place of , we may use our function
With this replacement, the diffeomorphism above becomes the diffeomorphism appearing in Example 2.12
Define and , so that we have -manifold isomorphisms
Applying the induced maps on -tangent bundles (see (2.1) in Section 2), we arrive at two complex -structures on which can be compared.
Claim 4.2.
The complex -structures on spanned and 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 and coincide up to a smooth rescaling over . Note restricts to a diffeomorphism of and recall on , on . Thus, we have diffeomorphisms
and it suffices to check that and agree up to a smooth rescaling. Indeed, since and are holomorphic in the usual sense away from , Proposition 2.13 shows that and are rescalings of (actually, from Lemma 3.2, we even know ). ∎
The preceding claim completes this section’s proof of Theorem 1.2 in the two-dimensional case. The desired local coordinates are provided by .
5 Existence of -holomorphic functions: general case
In this section, we extend the results of Section 3 to the general case. That is, we show that nontrivial -holomorphic functions exist near points on the singular locus of any complex -manifold. By Theorem 1.1, this amounts to the following.
Theorem 5.1.
Consider a complex -manifold with singular locus equipped with coordinates with complex -structure spanned by
where are linear combinations of with coefficients in smooth, complex-valued functions on vanishing to infinite order on . Then, there exists an open set containing and -holomorphic functions such that:
-
(i)
vanishes on and for ,
-
(ii)
for , 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
Let be a -invariant smooth function that vanishes to infinite order on . Then, the pushforward is a well-defined smooth function on that vanishes to infinite order on . ∎
Corollary 5.3.
Suppose are -invariant smooth functions vanishing infinite order on . Define (note is -invariant in the sense that ). Then, the pushforward is a well-defined smooth vector field on vanishing to infinite order on .
Lemma 5.4.
Let have coordinates . Suppose is a smooth, complex-valued function on vanishing to infinite order on . Then, and extend to smooth, complex-valued functions on vanishing to infinite order on . ∎
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 -holomorphic functions, there is no harm assuming the deformations terms to be -invariant. With this assumption, Lemma 3.2 and Corollary 5.3 together imply that the vector fields and have well-defined pushforwards by . Indeed,
where is a smooth vector field on vanishing to infinite order on for .
Because is a local diffeomorphism and bracket commutes with pushforward, , at least define a complex structure on . Going further, we can use Lemma 5.4 to write , where is another smooth vector field on vanishing to infinite order on . We claim that
(5.1) | |||||
(5.2) |
define an ordinary complex structure on the whole of (extending the aforementioned complex structure on ). Indeed, the deformation terms vanish on , 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 , hence everywhere by a continuity argument.
Now, by an application of the classical Newlander-Nirenberg theorem, there is an open neighbourhood of and smooth functions that are holomorphic for the complex structure defined by (5.1) and furthermore satisfy , for . In fact, again using the vanishing of the deformation terms along , the complex structure on defined by (5.1) is such that sits as a complex submanifold. Indeed, the inherited complex structure on is its standard one spanned by . 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 , we may even choose to vanish on . Putting for completes the proof of the result. ∎
6 The -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 -manifold with singular locus equipped with coordinates with complex -structure spanned by
where are smooth, complex vector fields on vanishing to infinite order on .
By Theorem 5.1, there is an open set containing and smooth functions satisfying:
-
(i)
for (-holomorphicity),
-
(ii)
vanishes on and for ,
-
(iii)
for .
As in Section 4, we decompose into real and imaginary parts . Since and vanish on , we may write and , where and are real-valued smooth functions. As in Section 4, by shrinking , we may enforce that is nowhere vanishing on and that:
-
(a)
is a diffeomorphism from onto an open set ,
-
(b)
and ,
-
(c)
where and where is .
As in Section 4, we argue that the pushforward by of the complex -structure on spanned by is independent of the deformation terms . Again, by Proposition 2.7, if suffices check this away from the singular locus . Using the fact that restricts to a self-diffeomorphism of , we have that defines a diffeomorphism of onto . Because the components of are holomorphic for the induced complex structure on , Proposition 2.13 implies that pushes forward the subbundle spanned by to the subbundle spanned by . Thus, the pushforward by of the subbundle spanned by is completely determined to be the pullback of the standard complex structure on by . 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 , played a crucial role in this article by enabling us to relate the nonelliptic operator to the standard complex partial derivative operator by way of the elementary, but perhaps somewhat mysterious, pushforward formula . 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 -Newlander-Nirenberg theorem for complex -manifolds [1].
Suppose that, more generally, we wish to relate the vector field
to , for a positive integer. More generally still, suppose we wish to relate
to , where is an entire function on whose restriction to 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 be a holomorphic function on so that is a holomorphic vector field on . We may decompose as
where the real vector fields and commute with one another. We may then define the complex-time flow of by
with the usual caveat that the flow need only be defined for sufficiently close to .
This flow is jointly holomorphic in and . One practical consequence of this joint holomorphicity is that, if is real on the -axis, so that coincides with on the -axis, then the complex-time flow of the holomorphic vector field may be obtained by analytic continuation in both variables of the real-time flow of the one-dimensional vector field .
Example 7.1.
The real-time flow of the one-dimensional vector field is given by . Accordingly, the complex-time flow of the holomorphic vector field is given by .
Example 7.2.
The real-time flow of the one-dimensional vector field is given by . Accordingly, the complex-time flow of the holomorphic vector field is given by .
Example 7.3.
For any integer , the real-time flow of the one-dimensional vector field is given by . Accordingly, the complex-time flow of the holomorphic vector field is given by .
Bearing in mind the above notations, we now give the main result of this section.
Proposition 7.4.
Let be a holomorphic function on that is real-valued on . Let be a sufficiently small open neighbourhood of that , is defined. Then, the vector field on is -related to the holomorphic vector field on .
Proof.
Decompose in the form , as above. Thus, by definition, . For fixed , taking sufficiently small, we have
Thus, the vector field is -related to . Similarly,
Thus, the vector field is -related to . The result follows by linearity. ∎
Corollary 7.5.
Let , . Then, is -related to .
Corollary 7.6.
Fix an integer . Let , . Then is -related to .
We remark that collapses to and defines a diffeomorphism . Note that the line is completely contained in the principal branch of . For purposes of clarification, sample plots of the real vector field in the decomposition are given below. Note all integral curves passing through the -axis are complete.



References
- [1] T. Barron and M. Francis. On automorphisms of complex -manifolds, 2023 (arXiv preprint).
- [2] M. Braverman, Y. Loizides, and Y. Song. Geometric quantization of -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 -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 -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.