Stemming from the Pythagorean Identity sin2z+cos2z=1 and Hörmander's L2-solution of the Cauchy-Riemann's equation ˉ∂u=f on C, this article demonstrates a corona-type principle which exists as a somewhat unexpected extension of the analytic Hilbert's Nullstellensatz on C to the quadratic Fock-Sobolev spaces on C.
Citation: Xiaofen Lv, Jie Xiao, Cheng Yuan. ˉ∂-equation look at analytic Hilbert's zero-locus theorem[J]. Electronic Research Archive, 2022, 30(1): 168-178. doi: 10.3934/era.2022009
[1] | Die Hu, Peng Jin, Xianhua Tang . The existence results for a class of generalized quasilinear Schrödinger equation with nonlocal term. Electronic Research Archive, 2022, 30(5): 1973-1998. doi: 10.3934/era.2022100 |
[2] | Shuai Yuan, Sitong Chen, Xianhua Tang . Normalized solutions for Choquard equations with general nonlinearities. Electronic Research Archive, 2020, 28(1): 291-309. doi: 10.3934/era.2020017 |
[3] | Shaban Khidr, Salomon Sambou . $ L^p $-theory for the $ {\partial }\overline{\partial} $-equation and isomorphisms results. Electronic Research Archive, 2025, 33(1): 68-86. doi: 10.3934/era.2025004 |
[4] | David Cheban, Zhenxin Liu . Averaging principle on infinite intervals for stochastic ordinary differential equations. Electronic Research Archive, 2021, 29(4): 2791-2817. doi: 10.3934/era.2021014 |
[5] |
Yuhui Chen, Ronghua Pan, Leilei Tong .
The sharp time decay rate of the isentropic Navier-Stokes system in |
[6] | Quanqing Li, Zhipeng Yang . Existence of normalized solutions for a Sobolev supercritical Schrödinger equation. Electronic Research Archive, 2024, 32(12): 6761-6771. doi: 10.3934/era.2024316 |
[7] | Jon Johnsen . Well-posed final value problems and Duhamel's formula for coercive Lax–Milgram operators. Electronic Research Archive, 2019, 27(0): 20-36. doi: 10.3934/era.2019008 |
[8] | Guaiqi Tian, Hongmin Suo, Yucheng An . Multiple positive solutions for a bi-nonlocal Kirchhoff-Schr$ \ddot{\mathrm{o}} $dinger-Poisson system with critical growth. Electronic Research Archive, 2022, 30(12): 4493-4506. doi: 10.3934/era.2022228 |
[9] | Massimo Grossi . On the number of critical points of solutions of semilinear elliptic equations. Electronic Research Archive, 2021, 29(6): 4215-4228. doi: 10.3934/era.2021080 |
[10] | Mingzhan Huang, Xiaohuan Yu, Shouzong Liu . Modeling and analysis of release strategies of sterile mosquitoes incorporating stage and sex structure of wild ones. Electronic Research Archive, 2023, 31(7): 3895-3914. doi: 10.3934/era.2023198 |
Stemming from the Pythagorean Identity sin2z+cos2z=1 and Hörmander's L2-solution of the Cauchy-Riemann's equation ˉ∂u=f on C, this article demonstrates a corona-type principle which exists as a somewhat unexpected extension of the analytic Hilbert's Nullstellensatz on C to the quadratic Fock-Sobolev spaces on C.
As one of the fundamentals of algebraic-complex geometry, the analytic Hilbert's Nullstellensatz (either theorem of zeros or zero-locus theorem) [1] on the finite complex plane C asserts that for finitely many analytic polynomials {pj}nj=1 without common zeros in C,
∃ finitely many analytic polynomials {qj}nj=1 such that n∑j=1pjqj=1. | (1) |
This celebrated principle has been improved and extended for more than a century; see e.g., Hermann [2], Masser-Wüstholz [3], Brownawell [4], Kollár [5], and Kwon-Neryanun-Trent [6] whose Lemma 1.4 especially indicates that an entire function Y is a polynomial on C if and only if lim|z|→∞|z|−m|Y(z)|=0 for some positive integer m.
Meanwhile, in complex trigonometry, the Pythagorean Identity on C states that
sin2z+cos2z=(eiz−e−iz2i)2+(eiz+e−iz2)2=1 ∀ z∈C, | (2) |
and sinz & cosz have no common zero as graphically shown below
![]() |
Although the entire functions sinz & cosz are not analytic polynomials, they can be appropriately approximated by analytic polynomials and satisfy:
{(sinz)g1(z)+(cosz)g2(z)=1;{g1(z)=sinz, g2(z)=cosz};supz∈C(|g1(z)|+|g2(z)|)e−|z|<∞;|sinz|+|cosz|≥1, |
where this last inequality is geometric due to the basic fact that the sum of the lengths of the adjacent & opposite sides BC & CA is not less than the length of the hypotenuse AB in the right triangle △ABC drawn below
![]() |
The previous two-fold observation actually inspires us to extend the analytic Hilbert's Nullstellensatz to some entire function spaces.
For α>0, let F2α be the Fock-Hilbert space of all L2(λα)-integrable entire functions (or analytic functions on C) with the inner product
⟨f,g⟩F2α=∫Cf(z)¯g(z)dλα(z) ∀ entire function pair {f,g}, |
where
dλα(z)=απ−1e−α|z|2dA(z)=απ−1e−α|z|2dxdy ∀ z=x+iy∈C. |
Moreover, for a nonnegative integer m, let F2α,m be the quadratic m-order Fock-Sobolev space of all entire functions obeying
‖f‖2F2α,m=∫C|zmf(z)|2e−α|z|2dA(z)<∞; |
Evidently,
α1<α2⟹F2α1,m⊆F2α2,m; |
see also Zhu's book [7] for more information.
Since all analytic polynomials are dense in F2α,m, as a somewhat unexpected variant of Eqs (1) and (2) we discover the following corona-type principle.
Theorem 1.1. Let
{α∈(0,∞);m,n∈{1,2,3,...};f1,...,fn∈F2α,1. |
If g is an entire function with
n∑j=1|fj|≥|g|m, | (3) |
then
∃ g1,...,gn∈F22mα,1 such that n∑j=1fjgj=g3m. | (4) |
Especially, if
n∑j=1|fj|≥1, | (5) |
then
∃ g1,...,gn∈F22mα,1 such that n∑j=1fjgj=1. | (6) |
For α∈(0,∞), let
C∋z↦{f(z)=∑∞n=0anzng(z)=∑∞n=0bnznbe two entire functions with their derivatives {f′(z),g′(z)}. |
Then some elementary calculations derive the following four formulae:
{∫Cf(z)¯g(z)dλα(z)=∑∞n=0α−nan¯bnn!;∫Czf(z)¯zg(z)dλα(z)=∑∞n=0α−n−1an¯bn(n+1)!;∫Cf′(z)¯g′(z)dλα(z)=∑∞n=1α1−nan¯bnn2(n−1)!=α∫Cf(z)¯g(z)(α|z|2−1)dλα(z);∫Czf(z)¯zg(z)dλα(z)=α−2∫Cf′(z)¯g′(z)dλα(z)+α−1∫Cf(z)¯g(z)dλα(z). |
Consequently,
F2α,1⊆F2α with ‖f‖2α,1=∫C|f′(z)|2dλα(z)⟹‖f‖2F2α,1=α−2‖f‖2α,1+α−1‖f‖2F2α. |
Moreover, a modification of Cho-Zhu's statement on [8, p. 2496] gives the following pointwise estimation
f∈F2α,1⟹|f(z)|≲‖f‖F2α,1(1+|z|)−1e2−1α|z|2 ∀ z∈C. | (1) |
Lemma 2.1. For
z=x+iy∈C;1<p<∞;p′=p(p−1)−1;C2c={all compactly-supported C2-functions: C→C};Ap(e−2ϕ)={all analytic functions in Lp(e−2ϕ)}, |
let ϕ:C→R & g:C→C be C2-smooth with
{0≤Δϕ(z)=4−1(∂2x+∂2y)ϕ(z)=∂zˉ∂zϕ(z)=∂ˉ∂ϕ(z);∂=∂z=2−1(∂∂x−i∂∂y);ˉ∂=∂ˉz=2−1(∂∂x+i∂∂y);ˉ∂∗2ϕg(z)=−e2ϕ(z)∂(g(z)e−2ϕ(z)). | (2) |
(i) Weighted (Lp,ˉ∂)-estimation - not only for a given function f on C there exists a weak solution u∈Lp(e−2ϕ) to ˉ∂u=f in the sense of
∫Cu¯ˉ∂∗2ϕge−2ϕdA=∫Cfˉge−2ϕdA ∀ g∈C2c | (3) |
if and only if
supg∈C2c|∫Cfˉge−2ϕdA|(∫C|ˉ∂∗2ϕg|p′e−2ϕdA)1p′<∞, | (4) |
but also Eq (4) holds for all f∈Lp((Δ(2ϕ)e2ϕ)−1) if and only if
∫C|g|p′(Δ(2ϕ))p′pe−2ϕdA≤∫C|ˉ∂∗2ϕg|p′e−2ϕdA ∀ g∈C2c. | (5) |
(ii) Uniqueness up to entire Lp(e−2ϕ)-functions - arbitrary two solutions in (i) differ by a function h∈Ap(e−2ϕ) with
∫C|h|pe−2ϕdA≤2p+1∫C|f|p(Δ(2ϕ)e2ϕ)−1dA. |
(iii) Weighted (Lp,ˉ∂)-Poincaré inequality - if u∈C1 satisfies (i) above and the Lp(e−2ϕ)-minimality below
∫C|u|pe−2ϕdA=infh∈Ap(e−2ϕ)∫C|u+h|pe−2ϕdA, | (6) |
then
∫C|u|pe−2ϕdA≤∫C|ˉ∂u|p(e2ϕΔ(2ϕ))−1dA, | (7) |
and consequently, if u∈C1 enjoys the case p=2 of (i) and the L2(e−2ϕ)-orthogonality
∫Cuˉhe−2ϕdA=0 ∀ h∈A2(e−2ϕ), | (8) |
then Eq (7) holds for p=2.
(iv) Weighted (L2,ˉ∂)-estimation is always available.
(v) Uniqueness - if
∃ ϵ∈(0,2) such that ∫C((1+|z|)ϵ−2e2ϕ(z))2dA(z)<∞, | (9) |
then the solution in (iv) is unique.
Proof. (i) This part is motivated by Berndtsson's [9, Theorems 2–3]. But, the argument comes from an adjustment of the case p=2 presented in Berndtsson's [10, Proposition 1.1].
Suppose that for a given function f on C there exists a weak solution u∈Lp(e−2ϕ) to ˉ∂u=f in the sense of Eq (3). Then the Hölder inequality derives
|∫Cfˉge−2ϕdA|=|∫Cu¯ˉ∂∗2ϕge−2ϕdA|≤(∫C|u|pe−2ϕdA)1p(∫C|ˉ∂∗2ϕg|p′e−2ϕdA)1p′ ∀ g∈C2c. |
Hence Eq (4) holds for the previous function f. Conversely, if Eq (4) is valid for a given function f on C, then
Sϕ={ˉ∂∗2ϕg: g∈C2c} |
is a subspace of Lp′(e−2ϕ), and hence the given function f induces the following bounded antilinear functional on Sϕ:
Lf(ˉ∂∗2ϕg)=∫Cfˉge−2ϕdA. |
This, along with the Hahn-Banach extension theorem, ensures an extension of Lf from Sϕ to Lp′(e−2ϕ). Consequently, the Riesz-type representation theorem for the dual of Lp′(e−2ϕ) produces a function
u∈[Lp′(e−2ϕ)]∗=Lp(e−2ϕ) |
such that
Lf(G)=∫Cu¯Ge−2ϕdA ∀ G∈Lp′(e−2ϕ). |
Upon taking G=ˉ∂∗2ϕg, we find that the last function u∈Lp(e−2ϕ) is a weak solution to ˉ∂u=f in the sense of Eq (3).
Moreover, if Eq (4) holds for all f∈Lp((Δ(2ϕ)e2ϕ)−1), then an application of the duality
[Lp(e−2ϕ)]∗=Lp′(e−2ϕ) |
under the pairing
⟨f,g⟩2ϕ=∫Cfˉge−2ϕdA=∫C((Δ(2ϕ))−1pf)((Δ(2ϕ))1pˉg)e−2ϕdA |
derives Eq (5). Evidently, if Eq (5) holds, then ⟨f,g⟩2ϕ deduces that Eq (4) is valid for all f∈Lp((Δ(2ϕ)e2ϕ)−1).
As an aside of the preceding demonstration, we achieve that for any f∈Lp((Δ(2ϕ)e2ϕ)−1) there exists a weak solution u∈Lp(e−2ϕ) of ˉ∂u=f with
∫C|u|pe−2ϕdA≤∫C|f|p(Δ(2ϕ))−1e−2ϕdA |
if and only if Eq (5) holds.
(ii) This follows from the fact that ˉ∂u=0 if and only if u is analytic.
(iii) This comes from a modification of the argument for Berndtsson's [10, Corollary 1.4]. Indeed, without loss of generality, we may assume
∫C|ˉ∂u|p(e2ϕΔ(2ϕ))−1dA<∞. |
Now, the verification of (i) ensures that
ˉ∂v=ˉ∂u∈Lp((Δ(2ϕ)e2ϕ)−1) |
has a weak solution v enjoying the inequality
∫C|v|pe−2ϕdA≤∫C|ˉ∂u|p(Δ(2ϕ))−1e−2ϕdA. | (10) |
Note that (ii) produces a function h†∈Ap(e−2ϕ) such that v=u+h†. So Eqs (6) & (10) imply
∫C|u|pe−2ϕdA≤∫C|u+h†|pe−2ϕdA≤∫C|ˉ∂u|p(Δ(2ϕ))−1e−2ϕdA, |
as desired in Eq (7).
Especially, if Eq (8) is valid, then a combination of
ˉ∂(v+h)=ˉ∂u ∀ h∈A2ϕ |
and the closedness of A2(e−2ϕ) in L2(e−2ϕ) yields a function h‡∈A2(e−2ϕ) such that
{∫C|v+h‡|2e−2ϕdA=infh∈A2ϕ∫C|v+h|2e−2ϕdA≤∫C|v|2e−2ϕdA;∫C(v+h‡−u)ˉhe−2ϕdA=0 ∀ h∈A2(e−2ϕ). | (11) |
Upon noticing ˉ∂(v+h‡−u)=0, we obtain
v+h‡−u∈(A2(e−2ϕ))∩(A2(e−2ϕ))⊥={0} & v+h‡=u. |
As a consequence, u is the L2(e−2ϕ)-minimal solution to ˉ∂v=ˉ∂u. Thus, the weighted (L2,ˉ∂)-Poincaré inequality follows from the first inequality of Eq (11) and the case p=2 of Eq (10).
Here it is appropriate to mention that as shown in [10, Theorem 3.3] the case p=2 of Eq (10) can be used to establish the following Brunn-Minkowski-type concavity: if D is a convex open subset of the (n+1)-dimesnional Euclidean space Rn×(−∞,∞) and Dt={x:(x,t)∈D} then the Lebesgue measure Mn(Dt) satisfies ∂2tlogMn(Dt)≤0 - i.e., - the function t↦logMn(Dt) is concave - in particular - so is t↦logA(Dt)=logM2(Dt).
(iv) This is a minor variant of [11, Theorem 1.1] - the well-known Hörmander L2-estimate for the ˉ∂-equation presented in [12]. In fact, given f∈L2((e2ϕΔ(2ϕ))−1) the basic identity (cf. [10, Proposition 1.2])
∫C|g|2(Δ(2ϕ))e−2ϕdA+∫C|ˉ∂g|2e−2ϕdA=∫C|ˉ∂∗2ϕg|2e−2ϕdA ∀ g∈C2c, |
ensures the second iff-condition of (i) with p=2
∫C|g|2(Δ(2ϕ))e−2ϕdA≤∫C|ˉ∂∗2ϕg|2e−2ϕdA ∀ g∈C2c, |
thereby reaching the existence of a weak solution u to ˉ∂u=f with
∫C|u|2e−2ϕdA≤∫C|f|2(Δ(2ϕ))−1e−2ϕdA. |
(v) Such a uniqueness is newly induced by Eq (1). Yet, its proof is similar to the argument for Hedenmalm's curvature-orientated uniqueness of the ˉ∂-equation in [11, Theorem 1.4]. As a matter of fact, if u1&u2 are two solutions in (iv), then u1−u2 is an entire function on C due to (ii), and hence log|u1−u2| is subharmonic on C. This, plus Eq (2), deduces
Δlog(|u1−u2|eϕ)=Δlog|u1−u2|+Δϕ>0, |
and so that
|u1−u2|eϕ=exp(log(|u1−u2|eϕ)) |
is subharmonic on C. Now, for any
(z0,r)∈C×(1+|z0|,∞), |
a combination of Eq (9), the mean-value-inequality for the subharmonic function |u1−u2|eϕ and the Cauchy-Schwarz inequality derives
|u1(z0)−u2(z0)|eϕ(z0)≤(πr2)−1∫|z−z0|<r|u1−u2|eϕdA≤(πr2)−1∫|z|<2r−1|u1−u2|eϕdA≤π−1(2r)−ϵ∫|z|<2r−1|u1(z)−u2(z)|eϕ(z)(1+|z|)ϵ−2dA(z)≤π−1(2r)−ϵ∫C|u1(z)−u2(z)|eϕ(z)(1+|z|)ϵ−2dA(z)≤π−1(2r)−ϵ(∫C|u1−u2|2dAe2ϕ)12(∫C((1+|z|)ϵ−2e2ϕ(z))2dA(z))12≤22π−1(2r)−ϵ(∫C|f|2dA(Δϕ)e2ϕ)12(∫C((1+|z|)ϵ−2e2ϕ(z))2dA(z))12. |
Letting r→∞ in the last estimation gives
{|u1(z0)−u2(z0)|eϕ(z0)=0;u1(z0)=u2(z0). |
Since z0 is arbitrary, the last equality ensures u1=u2 on C.
Argument for Theorem 1.1. Clearly, if g≡1 in (3)–(4), then (5)⟹(6) follows from Eq (3)⟹(4) which is verified as below.
Suppose that (3) is valid. Let
{φj=gm¯fjn∑l=1|fl|2∀ j∈{1,...,n};Hj,k=gmφjˉ∂φk∀ j,k∈{1,...,n}. |
If bj,k is a function solving pointwisely the ˉ∂-equation
ˉ∂bj,k=Hj,k, | (12) |
then each
gj=g2mφj+n∑k=1(bj,k−bk,j)fk | (13) |
is an entire function enjoying
n∑j=1gjfj=g2mn∑j=1fjφj+n∑k,j=1(bjk−bkj)fkfj=g3m, |
and hence the equation in (4) is met.
Thanks to the smoothness of f1,...,fn,g, Lemma 2.1(iv) with
ϕ(z)=2−1(2m−1)α|z|2 |
produces a function bj,k such that (12) holds pointwisely with
∫C|bj,k(z)|2e−(2m−1)α|z|2dA(z)≤2−1∫C|Hj,k(z)|2(e−(2m−1)α|z|2Δϕ(z))dA(z)=((2m−1)α)−1∫C|Hj,k(z)|2e−(2m−1)α|z|2dA(z). | (14) |
In order to achieve gj∈F22mα,1 in (4), in the sequel we employ (12)–(13) to prove
{∫C|Hj,k(z)|2e−(2m−1)α|z|2dA(z)<∞;∫C|zgj(z)|2e−2mα|z|2dA(z)<∞ | (15) |
⊳ It is easy to get
supz∈C|φj(z)|=supz∈C|gm(z)¯fj(z)∑nl=1|fl(z)|2|=supz∈C(|g(z)|m(∑nl=1|fl(z)|2)12)(|fj(z)|(∑nl=1|fl(z)|2)12)≲1. |
In the above and below, X≲Y stands for X≤cY for a positive constant c.
For the case m=1, we utilize Eqs (1) & (3) to derive
∫C|zg2m(z)φj(z)|2e−2mα|z|2dA(z)≲∫C|zg2(z)|2e−2α|z|2dA(z)≲‖g‖2F2α,1∫C|z|2|g(z)|2(e2−1α|z|21+|z|)2e−2α|z|2dA(z)≲‖g‖2F2α,1∫C|z|2|g(z)|2(1+|z|)−2e−α|z|2dA(z)≲‖g‖2F2α,1∫C|z|2|g(z)|2e−α|z|2dA(z)≲‖g‖4F2α,1≲n∑j=1‖fj‖4F2α,1<∞. |
For the case m>1, we utilize Eq (1) - the Hölder inequality - Eq (3) to derive
∫C|zg2m(z)φj(z)|2e−2mα|z|2dA(z)≲‖g‖4mF2α,1∫C(1+|z|)−4m|z|2dA(z)≲‖gm‖4F2α,1∫C(1+|z|)−2(2m−1)dA(z)≲n∑j=1‖fj‖4F2α,1<∞. |
In summary, we always have
∫C|zg2m(z)φj(z)|2e−2mα|z|2dA(z)≲n∑j=1‖fj‖4F2α,1<∞. | (16) |
Now, a straightforward computation gives
ˉ∂φj=gmn∑l=1fl(¯fl¯∂fj−¯fj¯∂fl)(n∑l=1|fl|2)2. |
As evaluated in [13,14], we have
|ˉ∂φj|2≲|g|2m(n∑l=1|fl|2)2n∑l=1|∂fl|2(n∑l=1|fl|2)4≲n∑l=1|∂fl|2n∑l=1|fl|2, |
thereby producing
|Hj,k|2=|gmφjˉ∂φk|2≲n∑l=1|∂fl|2. |
Clearly, we get
∫C|Hj,k(z)|2e−(2m−1)α|z|2dA(z)≲∫Cn∑l=1|∂fl(z)|2e−(2m−1)α|z|2dA(z)≲n∑l=1‖fl‖2F2α,1<∞, | (17) |
whence verifying the first inequality of (15).
⊳ According to Lemma 2.1(i), there exists bj,k classically resolving (12) with (14), and consequently, a combination of Eqs (1) & (17) yields
∫C|zbj,k(z)|2|fk(z)|2e−2mα|z|2dA(z)≲‖fk‖2F2α,1∫C|bj,k(z)|2|z|2(1+|z|)−2e−(2m−1)α|z|2dA(z)≲‖fk‖2F2α,1∫C|bj,k(z)|2e−(2m−1)α|z|2dA(z)≲‖fk‖2F2α,1∫C|Hj,k(z)|2e−(2m−1)α|z|2dA(z)≲‖fk‖2F2α,1n∑l=1‖fl‖2F2α,1<∞. | (18) |
Since the comparable constants in (18) are independent of {j,k}, the formula (13), along with (16) & (18), validates the second inequality of (15).
XL was supported by NNSF of China (#12171150; #11771139) & NSF of Zhejiang Province (LY20A010008); JX was supported by NSERC of Canada (#202979) & MUN's SBM-Fund (#214311); CY was supported by NNSF of China (#11501415).
The authors declare there is no conflict of interest.
[1] |
D. Hilbert, Über die vollen Invariantesysteme, Math. Ann., 42 (1893), 313–373. https://doi.org/10.1007/BF01444162 doi: 10.1007/BF01444162
![]() |
[2] |
G. Hermann, Die frage der endlich vielen schritte in der theorie der polynomideale, Math. Ann., 95 (1926), 736–788. https://doi.org/10.1007/BF01206635 doi: 10.1007/BF01206635
![]() |
[3] |
D. W. Masser, G. Wüstholz, Fields of large transcendence degree generated by values of elliptic functions, Invent. Math., 72 (1983), 407–464. https://doi.org/10.1007/BF01398396 doi: 10.1007/BF01398396
![]() |
[4] |
W. D. Brownawell, Bounds for the degrees in the Nullstellensatz, Ann. Math., 126 (1987), 577–591. https://doi.org/10.2307/1971361 doi: 10.2307/1971361
![]() |
[5] |
J. Kollár, Sharp effective Nullstellensatz, J. Amer. Math. Soc., 4 (1988), 963–975. https://doi.org/10.1090/S0894-0347-1988-0944576-7 doi: 10.1090/S0894-0347-1988-0944576-7
![]() |
[6] |
H. Kwon, A. Netyanun, T. Trent, An analytic approach to the degree bound in the Nullstellensatz, Proc. Amer. Math. Soc., 144 (2016), 1145–1152. https://doi.org/10.1090/proc/12781 doi: 10.1090/proc/12781
![]() |
[7] | K. Zhu, Analysis on Fock Spaces, Grad. Texts in Math., Vol. 263, Springer, New York, 2012. |
[8] |
H. R. Cho, K. Zhu, Fock-Sobolev spaces and their Carleson measures, J. Funct. Anal., 263 (2012), 2483–2506. https://doi.org/10.1016/j.jfa.2012.08.003 doi: 10.1016/j.jfa.2012.08.003
![]() |
[9] |
B. Berndtsson, Weighted estimates for ˉ∂ in domains in C, Duke Math. J., 66 (1992), 239–255. https://doi.org/10.1215/S0012-7094-92-06607-5 doi: 10.1215/S0012-7094-92-06607-5
![]() |
[10] | B. Berndtsson, An Introduction to Things ˉ∂, Analytic and Algebraic Geometry, McNeal, Amer. Math. Soc. Providence RI 2010. 7–76. https: //doi.org/10.1090/PCMS/017/02 |
[11] |
H. Hedenmalm, On Hörmander's solution of the ˉ∂-equation. I, Math. Z., 281 (2015), 349–355. https://doi.org/10.1007/s00209-015-1487-7 doi: 10.1007/s00209-015-1487-7
![]() |
[12] | L. Hörmander, An Introduction to Complex Analysis in Several Variables, North Holland, Amsterdam, 1973. |
[13] |
D. P. Banjade, Wolff's ideal theorem on Qp spaces, Rocky Mountain J. Math., 49 (2019), 2121–2133. https://doi.org/10.1216/RMJ-2019-49-7-2121 doi: 10.1216/RMJ-2019-49-7-2121
![]() |
[14] | J. Xiao, C. Yuan, Cauchy-Riemann ˉ∂-equations with some applications, Preprint (submitted), 2021. |
1. | Wenwan Yang, Cheng Yuan, Two variants of Hilbert’s Nullstellensatz, 2023, 51, 0092-7872, 1344, 10.1080/00927872.2022.2134411 | |
2. | Jie Xiao, Cheng Yuan, Hilbert’s Nullstellensatz for analytic trigonometric polynomials, 2022, 40, 07230869, 910, 10.1016/j.exmath.2022.09.005 |