<<

247A Notes on Lorentz spaces

Definition 1. For 1 ≤ p < ∞ and f : Rd → C we define ∗ 1/p (1) kfkLp ( d) := sup λ {x : |f(x)| > λ} weak R λ>0 and the weak Lp p d  ∗ L ( ) := f : kfk p d < ∞ . weak R Lweak(R ) p −p Equivalently, f ∈ Lweak if and only if |{x : |f(x)| > λ}| . λ . Warning. The quantity in (1) does not define a . This is the reason we append the asterisk to the usual norm notation. To make a side-by-side comparison with the usual Lp norm, we note that ZZ 1/p p−1 kfkLp = pλ dλ dx 0≤λ<|f(x)| Z ∞ dλ1/p = |{|f| > λ}| pλp 0 λ 1/p 1/p = p λ|{|f| > λ}| p dλ L ((0,∞), λ ) and, with the convention that p1/∞ = 1, ∗ 1/∞ 1/p kfkLp = p λ|{|f| > λ}| ∞ dλ . weak L ((0,∞), λ ) This suggests the following definition. Definition 2. For 1 ≤ p < ∞ and 1 ≤ q ≤ ∞ we define the Lorentz space Lp,q(Rd) as the space of measurable functions f for which ∗ 1/q 1/p (2) kfkLp,q := p λ|{|f| > λ}| q dλ < ∞. L ( λ ) p,p p p,∞ p From the discussion above, we see that L = L and L = Lweak. Again ∗ k · kLp,q is not a norm in general. Nevertheless, it is positively homogeneous: for all a ∈ C, ∗ −1 1/p ∗ (3) kafkLp,q = λ {|f| > |a| λ} Lq (dλ/λ) = |a| · kfkLp,q (strictly the case a = 0 should receive separate treatment). In lieu of the triangle , we have the following: ∗ 1/p kf + gkLp,q = λ {|f + g| > λ} Lq (dλ/λ) 1 1 1/p ≤ λ {|f| > 2 λ} + {|g| > 2 λ} Lq (dλ/λ) 1 1/p 1 1/p ≤ λ {|f| > 2 λ} Lq (dλ/λ) + λ {|g| > 2 λ} Lq (dλ/λ) by the of fractional powers and the in Lq(dλ/λ). Thus ∗ ∗ ∗ (4) kf + gkLp,q ≤ 2kfkLp,q + 2kgkLp,q .

Combining (3), (4), and the fact that kfkLp,q = 0 implies f ≡ 0 almost every- ∗ where, we see that k · kLp,q obeys the axioms of a quasi-norm. When p > 1, this quasi-norm is equivalent to an actual norm (see below). When p = 1 and q 6= 1, there cannot be a norm that is equivalent to our quasi-norm. However there is a that generates the same . In either case, we obtain a complete . ∗ ∗ Notice that (i) if |f| ≥ |g| then kfkLp,q ≥ kgkLp,q and (ii) The quasi-norms are ∗ ∗ rearrangement invariant, which is to say that kfkLp,q = kf ◦φkLp,q for any preserving bijection φ : Rd → Rd. 1 2

p,q P Proposition 3. Given f ∈ L , we write f = fm where

fm(x) := f(x)χ{x:2m≤|f(x)|<2m+1}. Then ∗ kfk p,q ≈p,q kfmk p d L Lx(R ) q `m(Z)

p,q1 p,q2 In particular, L ⊆ L whenever q1 ≤ q2. P m Proof. It suffices to consider f of the form f = 2 χEm with disjoint sets Em (cf. m m+1 Em = {2 ≤ |f| < 2 }). Now Z ∞ ∗ q q q/p dλ kfkLp,q = p λ |{|f| > λ}| 0 λ Z 2m  q/p X q X dλ = p λ |En| m−1 λ m 2 n≥m  1/p q X m X ≈ 2 |En| .

m n≥m To obtain a lower bound, we keep only the summand n = m; for an upper bound, we use the subadditivity of fractional powers. This yields

m 1/p ∗ X m 1/p (5) 2 |E | q kfk 2 |E | . m ` . Lp,q . n m q m≤n `m

m m 1/p p As k2 χEm kL = 2 |Em| , we have our desired lower bound. To obtain the upper bound, we use the triangle inequality in `q(Z): ∞ ∞ X −k m+k X −k m RHS(5) = 2 k2 χE kLp ≤ 2 k2 χE kLp m+k m q q `m k=0 `m k=0 This completes the proof of the upper bound.  Lemma 4. Given 1 ≤ q < ∞ and a finite set A ⊂ 2Z, X X q q X Aq ≤ A ≤ 2 max A ≤ 2q Aq A∈A where all sums are over A ∈ A. More generally, for any subset A of a geometric and any 0 < q < ∞, q X q X A ≈ A where the implicit constants depend on q and the step size of the geometric series. Proposition 5. For 1 < p < ∞ and 1 ≤ q ≤ ∞,  R ∗ ∗ (6) sup | fg| : kgkLp0,q0 ≤ 1 ≈ kfkLp,q . Indeed, LHS(6) defines a norm on Lp,q. Note that by (6), this norm is equivalent to our quasi-norm. Moreover, under this norm, Lp,q is a and when 0 0 q 6= ∞, the dual Banach space is Lp ,q , under the natural pairing. R Remark. When p = 1 (and q 6= 1), the LHS(6) is typically infinite; indeed, E |f| may well be infinite even for some set E of finite measure. In fact, there there cannot be a norm on Lp,q equivalent to our quasi-norm. For example, the impossibility of finding an equivalent norm for L1,∞(R) can be deduced by computing N ∗ N X −1 X −1 ∗ |x − n| ≈ N log(N) and |x − n| L1,∞ ≈ N. 1,∞ n=0 L n=0 3

Proof. Because the quasi-norm is positively homogeneous, we need only verify (6) in the case that f and g have quasi-norm comparable to one. We may also assume P n P m that f = 2 χFn and g = 2 χEm . By the normalization just mentioned, 0 X n 1/pq X m 1/p0 q (7) 2 |Fn| ≈ 1 ≈ 2 |Em| n m Combining the above with Lemma 4, we obtain q X X n 1/p X X n 1/pq (8) 2 A ≈ 2 |Fn| ≈ 1.

A∈2Z n:|Fn|≈A A∈2Z n:|Fn|≈A and similarly for g. Now we compute: Z X n m |fg| dx = 2 2 |Fn ∩ Em| n,m

X X n X m ≤ 2 · min(A, B) · 2

A,B∈2Z n:|Fn|∼A m:|Em|∼B

 1 1  0 X X n 1/p  A  0  B  p X m 1/p ≤ 2 A · min p , · 2 B . B A A,B∈2Z n:|Fn|∼A m:|Em|∼B Notice that this has the structure of a : two vectors (indexed over 2Z) with a sitting between them. Moreover, by Schur’s test, the matrix is a bounded on `q(2Z). Thus, Z X n 1/p X m 1/p0 |fg| dx . 2 A · 2 B ≈ 1 0 `q (A∈2Z) `q (B∈2Z) n:|Fn|∼A m:|Em|∼B by (8) and the corresponding statement for g. This completes proof of the . part P n p,q of (6). We turn now to the opposite inequality. Given f = 2 χFn ∈ L , we choose q−1  1  − 1 q−p X n p p0 X n(q−1) p g = 2 |Fn| |Fn| χFn = 2 |Fn| χFn . n n Then Z q−1 q q  1  1− 1  1    X n p n p0 X n p ∗ fg = 2 |Fn| 2 |Fn| = 2 |Fn| ≈ kfkLp,q ≈ 1. n n ∗ It remains to show that kgk 0 0 1. By Proposition 3, Lp ,q . q0/p0 q0 0 q−p ∗  X q X n(q−1) p kgk p0,q0 ≈ A |Fn| where n ∈ N(A) ⇔ 2 |Fn| ≈ A. L A∈2Z n∈N(A) Notice that for each A, the sum in n is over part of a geometric series; indeed,

p −n p(q−1) n ∈ N(A) ⇐⇒ |Fn| ≈ A q−p 2 q−p . Thus Lemma 4 applies and yields 0 ∗ q X q0 X q0/p0 X nq q/p kgkLp0,q0 ≈ A |Fn| ≈ 2 |Fn| ≈ 1. A∈2Z n∈N(A) n This provides the needed bound on g and so completes the proof of (6). The fact that LHS(6) is indeed a norm is a purely abstract statement about vector spaces and (separating) linear functionals. The proof that Lp,q is complete in this norm differs little from the usual Riesz–Fischer argument. p,q 1/p Let ` be a continuous linear functional on L . By definition, |`(χE)| . |E| and so the measure E 7→ `(χE) is absolutely continuous with respect to Lebesgue measure and so is represented by some locally L1 g. This is the Radon– Nikodym Theorem. By linearity this representation of the functional extends to 4

simple functions. Boundedness when tested against simple functions suffices to 0 0 show that g ∈ Lp ,q . When q 6= ∞, the simple functions are dense in Lp,q and so our linear functional admits the desired representation. When q = ∞ the simple functions are not dense. For example, one cannot approximate |x|−d/p ∈ Lp,∞(Rd) by simple functions. Indeed, inspired by the Banach limit linear functionals on `∞(Z) we can construct a non-trivial linear functional on Lp,∞ that vanishes on simple functions. Let L denote the of f ∈ Lp,∞ such that `(f) := lim |x|d/pf(x) exists. x→0 Notice that L contains the simple functions and that ` vanishes on these. By the p,q Hahn–Banach theorem, we can extended ` to a linear functional on all of L .  Definition 6. We say that a mapping T on (some class of) measurable functions is sublinear if it obeys

T (cf)(x) ≤ |c| T f(x) and T (f + g)(x) ≤ [T f](x) + [T g](x) for all c ∈ C and measurable functions f and g (in the domain of T ).

Linear maps are obviously sublinear. Moreover, if {Tt} is a family of linear maps then

[T f](x) := [Ttf](x) q Lt is sublinear. The case q = ∞ yields a kind of ‘maximal function’, while q = 2 gives a kind of ‘ function’.

Theorem 7 (Marcinkiewicz interpolation theorem). Fix 1 ≤ p0, p1, q0, q1 ≤ ∞ with p0 6= p1 and q0 6= q1. Let T be a sublinear operator that obeys

Z 0 1/q 1/pj (9) χE(x)[T χF ](x) dx . |E| j |F | j ∈ {0, 1}

uniformly for finite-measure sets E and F . Then for any 1 ≤ r ≤ ∞ and θ ∈ (0, 1), ∗ ∗ T f f Lqθ ,r . Lpθ ,r

where 1/pθ = (1 − θ)/p0 + θ/p1 and similarly, qθ = (1 − θ)/q0 + θ/q1. Remarks. 1. This form of the result is actually due to Hunt. The original version is Corollary 8 below. 2. Inequalities of the form (9) are known as restricted weak type estimates. Note Z 1/q0 1/p 1/p p χE[T χF ] dx . |E| |F | ⇔ T χF Lq,∞ . |F | ⇐ T f Lq,∞ . kfkL

as can be shown using Propositions 3 and 5. The rightmost inequality here is called a weak type estimate. At the top of the food chain sits the strong type estimate: kT fkLq . kfkLp . If pθ ≤ qθ we then we can choose r = qθ and so (using the nesting of Lorentz spaces) obtain a strong type estimate as the conclusion of the theorem. 3. The hypothesis pθ ≤ qθ is needed to obtain the strong type conclusion. Consider, for example,

−1/2 p  2p ,∞  f(x) 7→ x f(x) which maps L [0, ∞), dx → L p+2 [0, ∞), dx boundedly for all 2 ≤ p ≤ ∞. However

−1/p −1 − p+2 f(x) = x [log(x + x )] 2p

shows that T does not map Lp into L2p/(p+2) for any such p. 5

Proof of Theorem 7. By the relations among Lorentz spaces (cf. Proposi- tion 5), it suffices to show that Z ∗ ∗ g(x)[T f](x) dx . 1 whenever kfkLpθ ,r ≈ 1 ≈ kgk q0 ,r0 . L θ P m Moreover, we can take g to be of the form 2 χEm . We would like to take f of the same form, but this takes a little more justification. First by splitting a general f into real/imaginary parts and then each of these into its positive/negative parts, we see that it suffices to consider non-negative functions f. This also justifies taking g of the special form. Note that for g we can safely round up to the nearest power of two; however, since T need not have any monotonicity properties we are not able to do this for f. Now by using the binary expansion of the values of f(x) ≥ 0 at each point, we see that it is possible to write f as the sum of a functions of the form P n 1 1 2 χFn in such a way the summands are bounded pointwise by f, 2 f, 4 f, and so on. Since Lpθ ,qθ is a Banach space (specifically the triangle inequality holds) we can just sum the pieces back together. (A similar decomposition is possible under a quasi-norm, but a little cunning is required to avoid the summability being swamped by the constants from the triangle inequality.) P n P m Now we have reduced to considering f = 2 χFn and g = 2 χEm , let us compute:

Z 0 X n m 1/pj 1/q  |g(x)[T f](x)| dx . 2 2 min |Fn| |Em| j j∈{0,1} n,m   1 1   1 1 0 − 0 0 X X n 1/p p − p q q  X m 1/q . 2 A θ min A j θ B j θ 2 B θ . j∈{0,1} A,B∈2Z n:|Fn|∼A m:|Em|∼B Once again we recognize the structure of a bilinear form with vectors indexed over 2Z. With a little effort, we see that the matrix has the form 1 1 j−θ h 1 − 1 0 − 0 i  min A p1 p0 B q1 q0 j∈{0,1} r and so is bounded on ` (2Z) by Schur’s test. (It is essential here that p0 6= p1 and q0 6= q1.) On the other hand, by Lemma 4, r   r r X X n 1/pθ X n 1/pθ  ∗  2 A ≈ 2 |Fn| ≈ kfkLpθ ,r ≈ 1

A∈2Z n:|Fn|∼A n and similarly for g, though we use power r0. Putting these all together completes the proof. 

Corollary 8 (Marcinkiewicz interpolation theorem). Suppose 1 ≤ p0 < p1 ≤ ∞ and T is a sublinear operator that obeys

(10) T f kfk p0 and T f kfk p1 Lp0,∞ . L Lp1,∞ . L uniformly for measurable functions f. Then for any θ ∈ (0, 1),

kT fkLpθ . kfkLpθ where 1/pθ = (1 − θ)/p0 + θ/p1 and similarly, qθ = (1 − θ)/q0 + θ/q1.