Muckenhoupt weights

From HandWiki
Revision as of 18:43, 6 February 2024 by NBrush (talk | contribs) (update)
(diff) ← Older revision | Latest revision (diff) | Newer revision → (diff)

In mathematics, the class of Muckenhoupt weights Ap consists of those weights ω for which the Hardy–Littlewood maximal operator is bounded on Lp(). Specifically, we consider functions f on Rn and their associated maximal functions M( f ) defined as

[math]\displaystyle{ M(f)(x) = \sup_{r\gt 0} \frac{1}{r^n} \int_{B_r(x)} |f|, }[/math]

where Br(x) is the ball in Rn with radius r and center at x. Let 1 ≤ p < ∞, we wish to characterise the functions ω : Rn → [0, ∞) for which we have a bound

[math]\displaystyle{ \int |M(f)(x)|^p \, \omega(x) dx \leq C \int |f|^p \, \omega(x)\, dx, }[/math]

where C depends only on p and ω. This was first done by Benjamin Muckenhoupt.[1]

Definition

For a fixed 1 < p < ∞, we say that a weight ω : Rn → [0, ∞) belongs to Ap if ω is locally integrable and there is a constant C such that, for all balls B in Rn, we have

[math]\displaystyle{ \left(\frac{1}{|B|} \int_B \omega(x) \, dx \right)\left(\frac{1}{|B|} \int_B \omega(x)^{-\frac{q}{p}} \, dx \right)^\frac{p}{q} \leq C \lt \infty, }[/math]

where |B| is the Lebesgue measure of B, and q is a real number such that: 1/p + 1/q = 1.

We say ω : Rn → [0, ∞) belongs to A1 if there exists some C such that

[math]\displaystyle{ \frac{1}{|B|} \int_B \omega(y) \, dy \leq C\omega(x), }[/math]

for all xB and all balls B.[2]

Equivalent characterizations

This following result is a fundamental result in the study of Muckenhoupt weights.

Theorem. A weight ω is in Ap if and only if any one of the following hold.[2]
(a) The Hardy–Littlewood maximal function is bounded on Lp(ω(x)dx), that is
[math]\displaystyle{ \int |M(f)(x)|^p \, \omega(x)\, dx \leq C \int |f|^p \, \omega(x)\, dx, }[/math]
for some C which only depends on p and the constant A in the above definition.
(b) There is a constant c such that for any locally integrable function f on Rn, and all balls B:
[math]\displaystyle{ (f_B)^p \leq \frac{c}{\omega(B)} \int_B f(x)^p \, \omega(x)\,dx, }[/math]
where:
[math]\displaystyle{ f_B = \frac{1}{|B|}\int_B f, \qquad \omega(B) = \int_B \omega(x)\,dx. }[/math]

Equivalently:

Theorem. Let 1 < p < ∞, then w = eφAp if and only if both of the following hold:
[math]\displaystyle{ \sup_{B}\frac{1}{|B|}\int_{B}e^{\varphi-\varphi_B}dx\lt \infty }[/math]
[math]\displaystyle{ \sup_{B}\frac{1}{|B|}\int_{B}e^{-\frac{\varphi-\varphi_B}{p-1}}dx\lt \infty. }[/math]

This equivalence can be verified by using Jensen's Inequality.

Reverse Hölder inequalities and A

The main tool in the proof of the above equivalence is the following result.[2] The following statements are equivalent

  1. ωAp for some 1 ≤ p < ∞.
  2. There exist 0 < δ, γ < 1 such that for all balls B and subsets EB, |E| ≤ γ |B| implies ω(E) ≤ δω(B).
  3. There exist 1 < q and c (both depending on ω) such that for all balls B we have:
[math]\displaystyle{ \frac{1}{|B|} \int_{B} \omega^q \leq \left(\frac{c}{|B|} \int_{B} \omega \right)^q. }[/math]

We call the inequality in the third formulation a reverse Hölder inequality as the reverse inequality follows for any non-negative function directly from Hölder's inequality. If any of the three equivalent conditions above hold we say ω belongs to A.

Weights and BMO

The definition of an Ap weight and the reverse Hölder inequality indicate that such a weight cannot degenerate or grow too quickly. This property can be phrased equivalently in terms of how much the logarithm of the weight oscillates:

(a) If wAp, (p ≥ 1), then log(w) ∈ BMO (i.e. log(w) has bounded mean oscillation).
(b) If f  ∈ BMO, then for sufficiently small δ > 0, we have eδfAp for some p ≥ 1.

This equivalence can be established by using the exponential characterization of weights above, Jensen's inequality, and the John–Nirenberg inequality.

Note that the smallness assumption on δ > 0 in part (b) is necessary for the result to be true, as −log|x| ∈ BMO, but:

[math]\displaystyle{ e^{-\log|x|}=\frac{1}{e^{\log|x|}} = \frac{1}{|x|} }[/math]

is not in any Ap.

Further properties

Here we list a few miscellaneous properties about weights, some of which can be verified from using the definitions, others are nontrivial results:

[math]\displaystyle{ A_1 \subseteq A_p \subseteq A_\infty, \qquad 1\leq p\leq\infty. }[/math]
[math]\displaystyle{ A_\infty = \bigcup_{p\lt \infty}A_p. }[/math]
If wAp, then wdx defines a doubling measure: for any ball B, if 2B is the ball of twice the radius, then w(2B) ≤ Cw(B) where C > 1 is a constant depending on w.
If wAp, then there is δ > 1 such that wδAp.
If wA, then there is δ > 0 and weights [math]\displaystyle{ w_1,w_2\in A_1 }[/math] such that [math]\displaystyle{ w=w_1 w_2^{-\delta} }[/math].[3]

Boundedness of singular integrals

It is not only the Hardy–Littlewood maximal operator that is bounded on these weighted Lp spaces. In fact, any Calderón-Zygmund singular integral operator is also bounded on these spaces.[4] Let us describe a simpler version of this here.[2] Suppose we have an operator T which is bounded on L2(dx), so we have

[math]\displaystyle{ \forall f \in C^{\infty}_c : \qquad \|T(f)\|_{L^2} \leq C\|f\|_{L^2}. }[/math]

Suppose also that we can realise T as convolution against a kernel K in the following sense: if f , g are smooth with disjoint support, then:

[math]\displaystyle{ \int g(x) T(f)(x) \, dx = \iint g(x) K(x-y) f(y) \, dy\,dx. }[/math]

Finally we assume a size and smoothness condition on the kernel K:

[math]\displaystyle{ \forall x \neq 0, \forall |\alpha| \leq 1 : \qquad \left |\partial^{\alpha} K \right | \leq C |x|^{-n-\alpha}. }[/math]

Then, for each 1 < p < ∞ and ωAp, T is a bounded operator on Lp(ω(x)dx). That is, we have the estimate

[math]\displaystyle{ \int |T(f)(x)|^p \, \omega(x)\,dx \leq C \int |f(x)|^p \, \omega(x)\, dx, }[/math]

for all f for which the right-hand side is finite.

A converse result

If, in addition to the three conditions above, we assume the non-degeneracy condition on the kernel K: For a fixed unit vector u0

[math]\displaystyle{ |K(x)| \geq a |x|^{-n} }[/math]

whenever [math]\displaystyle{ x = t \dot u_0 }[/math] with −∞ < t < ∞, then we have a converse. If we know

[math]\displaystyle{ \int |T(f)(x)|^p \, \omega(x)\,dx \leq C \int |f(x)|^p \, \omega(x)\, dx, }[/math]

for some fixed 1 < p < ∞ and some ω, then ωAp.[2]

Weights and quasiconformal mappings

For K > 1, a K-quasiconformal mapping is a homeomorphism f  : RnRn such that

[math]\displaystyle{ f\in W^{1,2}_{loc}(\mathbf{R}^n), \quad \text{ and } \quad \frac{\|Df(x)\|^n}{J(f,x)}\leq K, }[/math]

where Df (x) is the derivative of f at x and J( f , x) = det(Df (x)) is the Jacobian.

A theorem of Gehring[5] states that for all K-quasiconformal functions f  : RnRn, we have J( f , x) ∈ Ap, where p depends on K.

Harmonic measure

If you have a simply connected domain Ω ⊆ C, we say its boundary curve Γ = ∂Ω is K-chord-arc if for any two points z, w in Γ there is a curve γ ⊆ Γ connecting z and w whose length is no more than K|zw|. For a domain with such a boundary and for any z0 in Ω, the harmonic measure w( ⋅ ) = w(z0, Ω, ⋅) is absolutely continuous with respect to one-dimensional Hausdorff measure and its Radon–Nikodym derivative is in A.[6] (Note that in this case, one needs to adapt the definition of weights to the case where the underlying measure is one-dimensional Hausdorff measure).

References

  • Garnett, John (2007). Bounded Analytic Functions. Springer. 
  1. Muckenhoupt, Benjamin (1972). "Weighted norm inequalities for the Hardy maximal function". Transactions of the American Mathematical Society 165: 207–226. doi:10.1090/S0002-9947-1972-0293384-6. 
  2. 2.0 2.1 2.2 2.3 2.4 Stein, Elias (1993). "5". Harmonic Analysis. Princeton University Press. 
  3. Jones, Peter W. (1980). "Factorization of Ap weights". Ann. of Math.. 2 111 (3): 511–530. doi:10.2307/1971107. 
  4. Grafakos, Loukas (2004). "9". Classical and Modern Fourier Analysis. New Jersey: Pearson Education, Inc.. 
  5. Gehring, F. W. (1973). "The Lp-integrability of the partial derivatives of a quasiconformal mapping". Acta Math. 130: 265–277. doi:10.1007/BF02392268. 
  6. Garnett, John; Marshall, Donald (2008). Harmonic Measure. Cambridge University Press.