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

A Counterexample to CkC^{k}-regularity for the Newlander-Nirenberg Theorem

Yao, Liding
University of Wisconsin-Madison
Abstract

We give an example of CkC^{k}-integrable almost complex structure that does not admit a corresponding Ck+1C^{k+1}-complex coordinate system.

The celebrated Newlander-Nirenberg theorem [5] states that given an integrable almost complex structure, it is locally induced by some complex coordinate system.

Malgrange [4] proved the existence of such complex coordinate chart when the almost complex structure is not smooth, and he obtained the following sharp Hölder regularity for this chart:

Theorem 1 (Sharp Newlander-Nirenberg).

Let k+k\in\mathbb{Z}_{+} and 0<α<10<\alpha<1, let M2nM^{2n} be a Ck+1,αC^{k+1,\alpha}-manifold endowed with a Ck,αC^{k,\alpha}-almost complex structure J:TMTMJ:TM\to TM. If JJ is integrable, then for any pMp\in M there is a Ck+1,αC^{k+1,\alpha}-complex coordinate chart (w1,,wn):Un(w^{1},\dots,w^{n}):U\to\mathbb{C}^{n} near pp such that Jwj=iwjJ\frac{\partial}{\partial w^{j}}=i\frac{\partial}{\partial w^{j}} for j=1,,nj=1,\dots,n.

There are several equivalent characterizations for integrability. One of which is the vanishing of the Nijenhuis tensor NJ(X,Y):=[JX,JY]J[JX,Y]J[X,JY][X,Y]N_{J}(X,Y):=[JX,JY]-J[JX,Y]-J[X,JY]-[X,Y].

Our main theorem is to show that this is not true when α{0,1}\alpha\in\{0,1\}:

Theorem 2.

Let k,n+k,n\in\mathbb{Z}_{+}. There is a CkC^{k}-integrable almost complex structure JJ on 2n\mathbb{R}^{2n}, such that there is no Ck,1C^{k,1}-complex coordinate chart w:U2nnw:U\subset\mathbb{R}^{2n}\to\mathbb{C}^{n} near 0 satisfying Jwj=iwjJ\frac{\partial}{\partial w^{j}}=i\frac{\partial}{\partial w^{j}} for j=1,,nj=1,\dots,n.

For convenience we use the viewpoint of eigenbundle of JJ: Set 𝒱n=p{vTp2n:Jpv=iv}\mathcal{V}_{n}=\coprod_{p}\{v\in\mathbb{C}T_{p}\mathbb{R}^{2n}:J_{p}v=iv\}. So Jwj=iwjJ\frac{\partial}{\partial w^{j}}=i\frac{\partial}{\partial w^{j}} for all jj iff dw¯1,,dw¯nd\bar{w}^{1},\dots,d\bar{w}^{n} spans 𝒱n|UT2n|U\mathcal{V}_{n}^{\bot}|_{U}\leq\mathbb{C}T^{*}\mathbb{R}^{2n}|_{U}. And JJ is integrable if and only if X,YΓ(𝒱n)[X,Y]Γ(𝒱n)X,Y\in\Gamma(\mathcal{V}_{n})\Rightarrow[X,Y]\in\Gamma(\mathcal{V}_{n}) for all complex vector fields X,YX,Y. See [1] Chapter 1 for details.

First we can restrict our focus to the 1-dimensional case:

Proof of 1-dim \Rightarrow n-dim.

Suppose J1J_{1} is a CkC^{k}-almost complex structure on z11\mathbb{C}^{1}_{z^{1}} (not compatible with the standard complex structure), such that near 0, there is no Ck,1C^{k,1}-complex coordinate φ\varphi satisfying 𝒱1=Spandφ¯\mathcal{V}_{1}^{\bot}=\operatorname{Span}d\bar{\varphi} in the domain. Here 𝒱1\mathcal{V}_{1}^{\bot} is the dual eigenbundle of J1J_{1}.

Denote θ=θ(z1)\theta=\theta(z^{1}) as a CkC^{k} 1-form on z11\mathbb{C}^{1}_{z^{1}} that spans 𝒱1\mathcal{V}_{1}^{\bot}.

Consider 2n(z1,,zn)n\mathbb{R}^{2n}\simeq\mathbb{C}^{n}_{(z^{1},\dots,z^{n})}. We identify θ\theta as the CkC^{k} 1-form on 2n\mathbb{R}^{2n}. Take an nn-dim almost complex structure on 2n\mathbb{R}^{2n} such that the dual of eigenbundle 𝒱n\mathcal{V}_{n}^{\bot} is spanned by θ,dz¯2,,dz¯n\theta,d\bar{z}^{2},\dots,d\bar{z}^{n}.

In other words, 𝒱n\mathcal{V}_{n} is the “tensor” of 𝒱1\mathcal{V}_{1} with the standard complex structure of (z2,,zn)n1\mathbb{C}^{n-1}_{(z^{2},\dots,z^{n})}.

If w=(w1,,wn)w=(w^{1},\dots,w^{n}) is a corresponding Ck,1C^{k,1}-complex chart for 𝒱n\mathcal{V}_{n} near 0, then there is a 1j0n1\leq j_{0}\leq n such that dw¯j00(moddz¯2,,dz¯n)d\bar{w}^{j_{0}}\not\equiv 0\pmod{d\bar{z}^{2},\dots,d\bar{z}^{n}} near 0. In other words, we have linear combinations dw¯j0=λ1θ+λ2dz¯2++λndz¯nd\bar{w}^{j_{0}}=\lambda_{1}\theta+\lambda_{2}d\bar{z}^{2}+\dots+\lambda_{n}d\bar{z}^{n} for some non-vanishing function λ1(z1,,zn)\lambda_{1}(z^{1},\dots,z^{n}) near z=0z=0.

Therefore wj0(,0n1)w^{j_{0}}(\cdot,0^{n-1}) is a complex coordinate chart defined near z1=02z^{1}=0\in\mathbb{R}^{2} whose differential spans 𝒱n|z12𝒱1\mathcal{V}_{n}^{\bot}|_{\mathbb{R}^{2}_{z^{1}}}\cong\mathcal{V}_{1}^{\bot} near 0. By our assumption on 𝒱1\mathcal{V}_{1}, we have wj0(,0)Ck,1w^{j_{0}}(\cdot,0)\notin C^{k,1}. So wj0Ck,1w^{j_{0}}\notin C^{k,1}, which means wCk,1w\notin C^{k,1}. ∎

Now we focus on the one-dimensional case. Note that a 1-dim structure is automatically integrable.

Fix k1k\geq 1. Define an almost complex structure by setting its eigenbundle 𝒱1T2\mathcal{V}_{1}\leq\mathbb{C}T\mathbb{R}^{2} equals to the span of z+a(z)z¯\frac{\partial}{\partial z}+a(z)\frac{\partial}{\partial\bar{z}}, where aCk(2;)a\in C^{k}(\mathbb{R}^{2};\mathbb{C}) has compact support that satisfies the following:

  1. (i)

    aCloc(2\{0};)a\in C^{\infty}_{\mathrm{loc}}(\mathbb{R}^{2}\backslash\{0\};\mathbb{C});

  2. (ii)

    z1z¯aCk1,1\partial_{z}^{-1}\partial_{\bar{z}}a\notin C^{k-1,1} near 0;

  3. (iii)

    zaCk+1(2;)za\in C^{k+1}(\mathbb{R}^{2};\mathbb{C}) and z1aCk1(2;)z^{-1}a\in C^{k-1}(\mathbb{R}^{2};\mathbb{C}) (which implies a(z)=o(|z|)a(z)=o(|z|) as z0z\to 0);

  4. (iv)

    suppa𝔹2\operatorname{supp}a\subset\mathbb{B}^{2};

  5. (v)

    aC0<δ0\|a\|_{C^{0}}<\delta_{0} for some small enough δ0>0\delta_{0}>0 (take δ0=101\delta_{0}=10^{-1} will be ok).

Here we take z1\partial_{z}^{-1} to be the conjugated Cauchy-Green operator on the unit disk111Throughout the paper, 𝔹2=B2(0,1)\mathbb{B}^{2}=B^{2}(0,1) refers to the unit disk in 1\mathbb{C}^{1}.:

z1ϕ(z)=z,𝔹21ϕ(z):=1π𝔹2ϕ(ξ+iη)dξdηz¯ξ+iη.\displaystyle\partial_{z}^{-1}\phi(z)=\partial_{z,\mathbb{B}^{2}}^{-1}\phi(z):=\frac{1}{\pi}\int_{\mathbb{B}^{2}}\frac{\phi(\xi+i\eta)d\xi d\eta}{\bar{z}-\xi+i\eta}.

We use notation z1\partial_{z}^{-1} because it is an right inverse of z\partial_{z}. And z1:Cm,β(𝔹2;)Cm+1,β(𝔹2;)\partial_{z}^{-1}:C^{m,\beta}(\mathbb{B}^{2};\mathbb{C})\to C^{m+1,\beta}(\mathbb{B}^{2};\mathbb{C}) is bounded linear for all m0m\in\mathbb{Z}_{\geq 0}, 0<β<10<\beta<1. See [8] theorem 1.32 in section 8.1 (page 56), or [2] lemma 2.3.4 for example.

We can take suppa𝔹2\operatorname{supp}a\subset\mathbb{B}^{2} such that when |z|<12|z|<\frac{1}{2},

a(z):=1100z¯k+1z((log|z|)12)=1100z(z¯k+1(log|z|)12).\textstyle a(z):=\frac{1}{100}\bar{z}^{k+1}\partial_{z}\big{(}(-\log|z|)^{\frac{1}{2}}\big{)}=\frac{1}{100}\partial_{z}\big{(}\bar{z}^{k+1}(-\log|z|)^{\frac{1}{2}}\big{)}. (1)

Note that for this aa we have a(z)=O(|z|k(log|z|)12)=o(|z|k)a(z)=O\big{(}|z|^{k}(-\log|z|)^{-\frac{1}{2}}\big{)}=o(|z|^{k}).

Remark 3.

Roughly speaking, Property (i) Singsuppa={0}\operatorname{Singsupp}a=\{0\} says that the regularity of a(z)a(z) corresponds to the vanishing order of aa at 0. To some degree, by multiplying with a(z)a(z), a function gains some regularity at the origin.

We check Property (ii) that z1z¯aCk1,1\partial_{z}^{-1}\partial_{\bar{z}}a\notin C^{k-1,1} here.

Lemma 4.

Let a(z)a(z) be given by (1), and let χCc(12𝔹2)\chi\in C_{c}^{\infty}(\frac{1}{2}\mathbb{B}^{2}) satisfies χ1\chi\equiv 1 in a neighborhood of 0. Then z1(χaz¯)Ck1,1\partial_{z}^{-1}(\chi a_{\bar{z}})\notin C^{k-1,1} near z=0z=0.

Proof.

Denote b(z):=z¯k+1(log|z|)12b(z):=\bar{z}^{k+1}(-\log|z|)^{\frac{1}{2}}, so bC(12𝔹2\{0};)b\in C^{\infty}(\frac{1}{2}\mathbb{B}^{2}\backslash\{0\};\mathbb{C}), χa=1100χbz\chi a=\frac{1}{100}\chi b_{z} and χaz¯=1100χbzz¯\chi a_{\bar{z}}=\frac{1}{100}\chi b_{z\bar{z}}.

First we show that z1(χbzz¯)bz¯C(12𝔹2;)\partial_{z}^{-1}(\chi b_{z\bar{z}})-b_{\bar{z}}\in C^{\infty}(\frac{1}{2}\mathbb{B}^{2};\mathbb{C}). We write

z1(χbzz¯)bz¯=z1z(χbz¯)χbz¯(1χ)bz¯z1(χzbz¯).\partial_{z}^{-1}(\chi b_{z\bar{z}})-b_{\bar{z}}=\partial_{z}^{-1}\partial_{z}(\chi b_{\bar{z}})-\chi b_{\bar{z}}-(1-\chi)b_{\bar{z}}-\partial_{z}^{-1}(\chi_{z}b_{\bar{z}}).

Since zz1z(χbz¯)=z(χbz¯)\partial_{z}\partial_{z}^{-1}\partial_{z}(\chi b_{\bar{z}})=\partial_{z}(\chi b_{\bar{z}}), we know z1z(χbz¯)χbz¯\partial_{z}^{-1}\partial_{z}(\chi b_{\bar{z}})-\chi b_{\bar{z}} is anti-holomorphic. By Cauchy integral formula we get z1z(χbz¯)χbz¯C\partial_{z}^{-1}\partial_{z}(\chi b_{\bar{z}})-\chi b_{\bar{z}}\in C^{\infty}.

By assumption 0suppχz0\notin\operatorname{supp}\chi_{z}, 0supp(1χ)0\notin\operatorname{supp}(1-\chi) and bC(12𝔹2\{0};)b\in C^{\infty}(\frac{1}{2}\mathbb{B}^{2}\backslash\{0\};\mathbb{C}), we know χzbz¯Cc\chi_{z}b_{\bar{z}}\in C_{c}^{\infty} and (1χ)bz¯C(12𝔹2;)(1-\chi)b_{\bar{z}}\in C^{\infty}(\frac{1}{2}\mathbb{B}^{2};\mathbb{C}).

Note that when acting on functions supported in the unit disk, z1=1πz¯()\partial_{z}^{-1}=\frac{1}{\pi\bar{z}}\ast(\cdot) is a convolution operator with kernel 1πz¯L1\frac{1}{\pi\bar{z}}\in L^{1}, so z1(χzbz¯)=1πz¯(χzbz¯)C\partial_{z}^{-1}(\chi_{z}b_{\bar{z}})=\frac{1}{\pi\bar{z}}\ast(\chi_{z}b_{\bar{z}})\in C^{\infty}.

It remains to show bz¯Ck1,1b_{\bar{z}}\notin C^{k-1,1} near 0. Indeed one has

z¯k+1(z¯k+1(log|z|)12)=(k+1)!(log|z|)12+O(1), as z0,\partial_{\bar{z}}^{k+1}(\bar{z}^{k+1}(-\log|z|)^{\frac{1}{2}})=(k+1)!(-\log|z|)^{\frac{1}{2}}+O(1)\quad,\text{ as }z\to 0,

because by Leibniz rule

z¯kz¯(z¯k+1(log|z|)12)=j=0k+1(k+1j)z¯k+1j(zk+1)z¯j(log|z|)12\displaystyle\textstyle\partial_{\bar{z}}^{k}\partial_{\bar{z}}\big{(}\bar{z}^{k+1}(-\log|z|)^{\frac{1}{2}}\big{)}=\sum_{j=0}^{k+1}{k+1\choose j}\partial_{\bar{z}}^{k+1-j}(z^{k+1})\cdot\partial_{\bar{z}}^{j}(-\log|z|)^{\frac{1}{2}}
=\displaystyle= (k+1)!(log|z|)12+j=1k+1O(zj)O(zj(log|z|)12)=(k+1)!(log|z|)12+O((log|z|)12).\displaystyle\textstyle(k+1)!(-\log|z|)^{\frac{1}{2}}+\sum_{j=1}^{k+1}O(z^{j})O\big{(}z^{-j}(-\log|z|)^{-\frac{1}{2}}\big{)}=(k+1)!(-\log|z|)^{\frac{1}{2}}+O\big{(}(-\log|z|)^{-\frac{1}{2}}\big{)}.

Now assume w:U~2w:\tilde{U}\subset\mathbb{R}^{2}\to\mathbb{C} is a 1-dim C1C^{1}-complex coordinate chart defined near 0 that represents 𝒱1\mathcal{V}_{1}, then Spandw¯=Span(dz¯adz)|U~=𝒱1|U~\operatorname{Span}d\bar{w}=\operatorname{Span}(d\bar{z}-adz)|_{\tilde{U}}=\mathcal{V}_{1}^{\bot}|_{\tilde{U}}. So dw¯=w¯z¯dz¯+w¯zdz=w¯z¯(dz¯adz)d\bar{w}=\bar{w}_{\bar{z}}d\bar{z}+\bar{w}_{z}dz=\bar{w}_{\bar{z}}(d\bar{z}-adz), that is,

wz¯(z)+a¯(z)wz(z)=0,zU~.\displaystyle\frac{\partial w}{\partial\bar{z}}(z)+\bar{a}(z)\frac{\partial w}{\partial z}(z)=0,\qquad z\in{\tilde{U}}.
Remark 5.

It is worth noticing that wz+az¯\partial_{w}\neq\partial_{z}+a\partial_{\bar{z}}. Indeed w\partial_{w} is only a scalar multiple of z+az¯\partial_{z}+a\partial_{\bar{z}}.

Note that wz(0)0w_{z}(0)\neq 0 because (dz¯adz)|0=dz¯|0Spandw¯|0(d\bar{z}-adz)|_{0}=d\bar{z}|_{0}\in\operatorname{Span}d\bar{w}|_{0}. So by multiplying wz(0)1w_{z}(0)^{-1}, we can assume wz(0)=1w_{z}(0)=1 without loss of generality. Then f:=logzwf:=\log\partial_{z}w is a well-defined function in a smaller neighborhood UU~U\subset\tilde{U} of 0, which solves

fz¯(z)+a(z)¯fz(z)=a¯z(z)(=az¯(z)¯),zU.\frac{\partial f}{\partial\bar{z}}(z)+\overline{a(z)}\frac{\partial f}{\partial z}(z)=-\frac{\partial\bar{a}}{\partial z}(z)\quad\Big{(}=-\overline{\frac{\partial a}{\partial\bar{z}}(z)}\Big{)},\qquad z\in U. (2)

Property (v) indicates that the operator z¯+a¯z\partial_{\bar{z}}+\bar{a}\partial_{z} is a first order elliptic operator. Therefore we can consider a second order divergence form elliptic operator

L:=z(z¯+a¯z)L:=\partial_{z}(\partial_{\bar{z}}+\bar{a}\partial_{z})

whose coefficients are CkC^{k} globally and are CC^{\infty} outside the origin.

By the classical Schauder’s estimate (see [7] Theorem 4.2, or [3] Chapter 6 & 8), we have the following:

Lemma 6 (Schauder’s interior estimate).

Assume u,ψC0,1(𝔹2;)u,\psi\in C^{0,1}(\mathbb{B}^{2};\mathbb{C}) satisfy Lu=ψzLu=\psi_{z}. Let U2U\subset\mathbb{R}^{2} be a neighborhood of 0. The following hold:

  1. (a)

    If ψCloc(U\{0};)\psi\in C^{\infty}_{\mathrm{loc}}(U\backslash\{0\};\mathbb{C}), then uCloc(U\{0};)u\in C^{\infty}_{\mathrm{loc}}(U\backslash\{0\};\mathbb{C}).

  2. (b)

    If ψCk1,1(2;)\psi\in C^{k-1,1}(\mathbb{R}^{2};\mathbb{C}), then uClock,1ε(2;)u\in C^{k,1-\varepsilon}_{\mathrm{loc}}(\mathbb{R}^{2};\mathbb{C}) for all 0<ε<10<\varepsilon<1.

Our Theorem 2 for 1-dim case is done by the following proposition:

Proposition 7.

For any neighborhood Uz2U\subset\mathbb{R}^{2}_{z} of 0, there is no fCk1,1(U;)f\in C^{k-1,1}(U;\mathbb{C}) solving (2).

Proof.

Suppose there is a neighborhood U12𝔹2U\subset\frac{1}{2}\mathbb{B}^{2} of the origin, and a solution fCk1,1(U;)f\in C^{k-1,1}(U;\mathbb{C}) to (2).

Applying Lemma 6 (a) on (2) with u=fu=f and ψ=a¯z\psi=-\bar{a}_{z}, we know thatfC(U\{0};)f\in C^{\infty}(U\backslash\{0\};\mathbb{C}).

Take χCc(U,[0,1])\chi\in C_{c}^{\infty}(U,[0,1]) such that χ1\chi\equiv 1 in a smaller neighborhood of 0. Denote

g(z):=χ(z)f(z),h(z):=z¯g(z).g(z):=\chi(z)f(z),\qquad h(z):=\bar{z}g(z).

So g,hg,h are Ck1,1C^{k-1,1}-functions defined in 2\mathbb{R}^{2} that are also smooth away from 0, and satisfy the following:

gz¯+a¯gz=χz¯f+χza¯fχa¯z,g_{\bar{z}}+\bar{a}g_{z}=\chi_{\bar{z}}f+\chi_{z}\bar{a}f-\chi\bar{a}_{z}, (3)
hz¯+a¯hz=g+z¯(χz¯f+χza¯f)χz¯a¯z.\qquad h_{\bar{z}}+\bar{a}h_{z}=g+\bar{z}(\chi_{\bar{z}}f+\chi_{z}\bar{a}f)-\chi\bar{z}\bar{a}_{z}. (4)

By construction χ0\nabla\chi\equiv 0 holds in a neighborhood of 0, so χz¯f+χza¯fCc(U;)\chi_{\bar{z}}f+\chi_{z}\bar{a}f\in C_{c}^{\infty}(U;\mathbb{C}).

Under our assumption that fCk1,1(U;)f\in C^{k-1,1}(U;\mathbb{C}), then the key is to show that

  1. (I)

    hCck,1ε(2;)h\in C^{k,1-\varepsilon}_{c}(\mathbb{R}^{2};\mathbb{C}), ε(0,1)\forall\varepsilon\in(0,1). This implies:

  2. (II)

    a¯gzCck1,1ε(2;)\bar{a}g_{z}\in C^{k-1,1-\varepsilon}_{c}(\mathbb{R}^{2};\mathbb{C}), ε(0,1)\forall\varepsilon\in(0,1).

(I) By assumption z¯(χz¯f+χza¯f)Cc\bar{z}(\chi_{\bar{z}}f+\chi_{z}\bar{a}f)\in C^{\infty}_{c}, gCk1,1g\in C^{k-1,1}, and by Property (iii), χz¯a¯zCk(2;)\chi\bar{z}\bar{a}_{z}\in C^{k}(\mathbb{R}^{2};\mathbb{C}).

Applying Lemma 6 (b) to (4), with u=hu=h and ψ=g+z¯(χz¯f+χza¯f)χz¯a¯zCck(2;)\psi=g+\bar{z}(\chi_{\bar{z}}f+\chi_{z}\bar{a}f)-\chi\bar{z}\bar{a}_{z}\in C^{k}_{c}(\mathbb{R}^{2};\mathbb{C}), we get hCk,1ε(𝔹2;)h\in C^{k,1-\varepsilon}(\mathbb{B}^{2};\mathbb{C}), for all 0<ε<10<\varepsilon<1.

(II) When k2k\geq 2, we know z1aCk1z^{-1}a\in C^{k-1} for Property (iii). So for any ε(0,1)\varepsilon\in(0,1), one has a¯gzCk1,1ε\bar{a}g_{z}\in C^{k-1,1-\varepsilon} because

z,z¯(a¯gz)=a¯(zgz,z¯gz)+gzz,z¯a¯=z¯1a¯(hzz,hzz¯gz)+O(Ck2,1)Ck2,1ε.\nabla_{z,\bar{z}}(\bar{a}g_{z})=\bar{a}\cdot(\partial_{z}g_{z},\partial_{\bar{z}}g_{z})+g_{z}\nabla_{z,\bar{z}}\bar{a}=\bar{z}^{-1}\bar{a}\cdot(h_{zz},h_{z\bar{z}}-g_{z})+O(C^{k-2,1})\in C^{k-2,1-\varepsilon}.

When k=1k=1, for any z1,z22\{0}z_{1},z_{2}\in\mathbb{R}^{2}\backslash\{0\}, note that gz¯g_{\bar{z}} is smooth outside the origin, so a¯gzC0,1ε\bar{a}g_{z}\in C^{0,1-\varepsilon}:

|a¯gz(z1)a¯gz(z2)||a¯(z1)||gz(z1)gz(z2)|+|a¯(z1)a¯(z2)||gz(z2)|\displaystyle|\bar{a}g_{z}(z_{1})-\bar{a}g_{z}(z_{2})|\leq|\bar{a}(z_{1})||g_{z}(z_{1})-g_{z}(z_{2})|+|\bar{a}(z_{1})-\bar{a}(z_{2})||g_{z}(z_{2})|
\displaystyle\leq |z¯11a¯(z1)||z¯1gz(z1)z¯2gz(z2)|+|z¯11a¯(z1)||z¯1z¯2||gz(z2)|+|a¯(z1)a¯(z2)||gz(z2)|\displaystyle|\bar{z}_{1}^{-1}\bar{a}(z_{1})||\bar{z}_{1}g_{z}(z_{1})-\bar{z}_{2}g_{z}(z_{2})|+|\bar{z}_{1}^{-1}\bar{a}(z_{1})||\bar{z}_{1}-\bar{z}_{2}||g_{z}(z_{2})|+|\bar{a}(z_{1})-\bar{a}(z_{2})||g_{z}(z_{2})|
\displaystyle\leq z1aC0hC1ε|z1z2|1ε+z1aC0gC0,1|z1z2|+aC1gC0,1|z1z2|.\displaystyle\|z^{-1}a\|_{C^{0}}\|\nabla h\|_{C^{1-\varepsilon}}|z_{1}-z_{2}|^{1-\varepsilon}+\|z^{-1}a\|_{C^{0}}\|g\|_{C^{0,1}}|z_{1}-z_{2}|+\|a\|_{C^{1}}\|g\|_{C^{0,1}}|z_{1}-z_{2}|.

Here as a remark, |a¯gz(z1)a¯gz(z2)|a,g,h|z1z2|1ε|\bar{a}g_{z}(z_{1})-\bar{a}g_{z}(z_{2})|\lesssim_{a,g,h}|z_{1}-z_{2}|^{1-\varepsilon} still makes sense when z1z_{1} or z2=0z_{2}=0, though g¯z(0)\bar{g}_{z}(0) may not be defined. Indeed limz0a¯gz(z)\lim\limits_{z\to 0}\bar{a}g_{z}(z) exists because a¯gz\bar{a}g_{z} itself has bounded C0,1εC^{0,1-\varepsilon}-oscillation on 𝔹2\{0}\mathbb{B}^{2}\backslash\{0\}, and then the limit defines the value of a¯gz\bar{a}g_{z} at z=0z=0.

So for either case of kk, we have a¯gzCk1,1ε(2;)\bar{a}g_{z}\in C^{k-1,1-\varepsilon}(\mathbb{R}^{2};\mathbb{C}) for all ε(0,1)\varepsilon\in(0,1).

Based on consequence (II), for (3) we have

g=(gz¯1gz¯)+z¯1(χz¯f+χza¯f)z¯1(a¯gz)z¯1(χa¯z),in 𝔹2.g=(g-\partial_{\bar{z}}^{-1}g_{\bar{z}})+\partial_{\bar{z}}^{-1}(\chi_{\bar{z}}f+\chi_{z}\bar{a}f)-\partial_{\bar{z}}^{-1}(\bar{a}g_{z})-\partial_{\bar{z}}^{-1}(\chi\bar{a}_{z}),\qquad\text{in }\mathbb{B}^{2}. (5)

The right hand side of (5) consists of four terms, the first to the third are all CkC^{k}, while the last one is not Ck1,1C^{k-1,1}. We explain these as follows:

  • Since z(gz¯1gz¯)=0\partial_{z}(g-\partial_{\bar{z}}^{-1}g_{\bar{z}})=0 in 𝔹2\mathbb{B}^{2}, we know gz¯1gz¯g-\partial_{\bar{z}}^{-1}g_{\bar{z}} is anti-holomorphic, which is smooth in 𝔹2\mathbb{B}^{2}.

  • By assumption χz¯f+χza¯fCc(2;)\chi_{\bar{z}}f+\chi_{z}\bar{a}f\in C^{\infty}_{c}(\mathbb{R}^{2};\mathbb{C}), so z¯1(χz¯f+χza¯f)C(𝔹2;)\partial_{\bar{z}}^{-1}(\chi_{\bar{z}}f+\chi_{z}\bar{a}f)\in C^{\infty}(\mathbb{B}^{2};\mathbb{C}) as well.

  • By consequence (II) a¯gzCk1,1ε(2;)\bar{a}g_{z}\in C^{k-1,1-\varepsilon}(\mathbb{R}^{2};\mathbb{C}), for all ε(0,1)\varepsilon\in(0,1), so z¯1(a¯gz)Ck,1εCk(𝔹2;)\partial_{\bar{z}}^{-1}(\bar{a}g_{z})\in C^{k,1-\varepsilon}\subset C^{k}(\mathbb{B}^{2};\mathbb{C}).

  • However by Lemma 4, z¯1(χa¯z)Ck1,1\partial_{\bar{z}}^{-1}(\chi\bar{a}_{z})\notin C^{k-1,1} near 0.

Combining each term to the right hand side of (5), we know gCk1,1g\notin C^{k-1,1} near 0. Contradiction! ∎

Remark 8.

The key to the proof is the non-surjectivity of z:CkCk1\partial_{z}:C^{k}\to C^{k-1}, which we use to construct a function a(z)a(z) such that a(0)=0a(0)=0, Singsuppa={0}\operatorname{Singsupp}a=\{0\}, and z1z¯aCk\partial_{z}^{-1}\partial_{\bar{z}}a\notin C^{k}.

Remark 9.

For positive integer kk, Malgrange’s sharp estimate of Theorem 1 still holds for Zygmund spaces 𝒞k=Bk\mathscr{C}^{k}=B^{k}_{\infty\infty}, that is, given J𝒞kJ\in\mathscr{C}^{k}, there exists a 𝒞k+1\mathscr{C}^{k+1}-coordinate chart (w1,,wn)(w^{1},\dots,w^{n}), such that Jwj=iwjJ\frac{\partial}{\partial w^{j}}=i\frac{\partial}{\partial w^{j}}. One can also see [6] for details. A reason why our proof does not give a counterexample for Zygmund spaces, is that there does exist a f𝒞kf\in\mathscr{C}^{k} defined in a neighborhood of 0 that solves (2).

Acknowledgement

The author would like to express his appreciation to his advisor Prof. Brian T. Street for his help.

References

  • [1] Shiferaw Berhanu, Paulo D. Cordaro, and Jorge Hounie. An introduction to involutive structures, volume 6 of New Mathematical Monographs. Cambridge University Press, Cambridge, 2008.
  • [2] So-Chin Chen and Mei-Chi Shaw. Partial differential equations in several complex variables, volume 19 of AMS/IP Studies in Advanced Mathematics. American Mathematical Society, Providence, RI; International Press, Boston, MA, 2001.
  • [3] David Gilbarg and Neil S. Trudinger. Elliptic partial differential equations of second order. Classics in Mathematics. Springer-Verlag, Berlin, 2001. Reprint of the 1998 edition.
  • [4] B. Malgrange. Sur l’intégrabilité des structures presque-complexes. In Symposia Mathematica, Vol. II (INDAM, Rome, 1968), pages 289–296. Academic Press, London, 1969.
  • [5] A. Newlander and L. Nirenberg. Complex analytic coordinates in almost complex manifolds. Ann. of Math. (2), 65:391–404, 1957.
  • [6] Brian Street. Sharp regularity for the integrability of elliptic structures. J. Funct. Anal., 278(1):108290, 2020.
  • [7] Michael E. Taylor. Partial differential equations III. Nonlinear equations, volume 117 of Applied Mathematical Sciences. Springer, New York, second edition, 2011.
  • [8] I. N. Vekua. Generalized analytic functions. Pergamon Press, London-Paris-Frankfurt; Addison-Wesley Publishing Co., Inc., Reading, Mass., 1962.

University of Wisconsin-Madison, Department of Mathematics, 480 Lincoln Dr.,
Madison, WI, 53706
[email protected]

MSC 2010: 35J46 (Primary), 32Q60 and 35F05 (Secondary)