arXiv:1809.02449v2 [math.FA] 2 Oct 2020 atII xmlsadilsrtoso h theory the of illustrations and Examples III. Part topics other with Connections II. Part theorems the of proofs and Statements I. Part Date revisited inequalities Kakeya References Multilinear revisited inequalities 11. Brascamp–Lieb revisited inequalities 10. Classical 9. operators multilinear general more and Factorisation convexity and 8. Factorisation factorisation and interpolation 7. Complex 6. proof the of details proof case: the General of overview problem case: 5. optimisation General convex results a main case: 4. the Discrete of discussion and 3. Statement 2. .Introduction 1. t etme 08 eie t erayad3t Septembe 30th and February 5th revised 2018, September 4th : arysfnaetltermo atrsto foperator pr of and factorisation on operators, theorem of r fundamental We factorisation Maurey’s consideration. f of under of theory inequality properties Nikisin–Stein norm factorisation the pointwise operators. to certain associated positive of of terms means in geometric weighted pointwise h hoyivle ovxotmsto n iia theory of minimax dual and the optimisation concerning tions convex involves theory the Abstract. nldn omsWinyieulte,Bacm–ibin concret Brascamp–Lieb inequalities. of inequalities, context the Loomis–Whitney in theory including the of ramifications the cuss ecnie h oncin fteter ihteter fi of theory the with theory the of connections the consider We NHN ABR,TM .H S. TIMO CARBERY, ANTHONY UTLNA ULT N ATRSTO FOR FACTORISATION AND DUALITY MULTILINEAR eiiit h td fadaiyter hc ple onor to applies which theory duality a of study the initiate We RSAPLE-YEINEQUALITIES BRASCAMP–LIEB-TYPE L ∞ n h oiaHwt hoyo ntl diiemeasures additive finitely of theory Yosida–Hewitt the and , NIE N STEF AND ANNINEN ¨ Contents 1 nto pcswihaenaturally are which spaces unction sn ul utlna eso of version multilinear fully a esent aiiso emti inequalities, geometric of families e qaiisadmliierKakeya multilinear and equalities through s troaino prtr.W dis- We operators. of nterpolation lt u hoyt h Maurey– the to theory our elate h hoyfid t expression its finds theory The ucinlaayi considera- functional-analytic , 2020. r NIG VALDIMARSSON INGI AN ´ L 1 h eeomn of development The . nqaiisfor inequalities m . 56 53 48 43 43 41 38 37 37 24 21 18 12 12 2 2 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON

1. Introduction In this paper we introduce and develop a general functional-analytic principle which gives a unifying framework for a range of multilinear phenomena that have recently arisen in a number of areas of mathematical analysis. We shall be mainly concerned with norm inequalities for pointwise weighted geometric means d αj (Tjfj (x)) j=1 Y of positive linear operators T defined on suitable spaces, where α 0 and d α = 1. Before j j ≥ j=1 j we describe the scope of our work in this paper, and to set the scene for our study, we briefly visit the analogous territory in the linear setting (d = 1) in order to help provideP a context for what we are aiming to achieve. Throughout the whole paper we shall be dealing with real-valued rather than complex-valued functions.

1.1. The linear setting. Let X and Y be measure spaces and let T : Lp(Y ) Lq(X) be a bounded linear operator, that is, it satisfies → (1) Tf A f k kq ≤ k kp for all f Lp(Y ), for some A> 0. Here, 1 p and 0

(2) (Tf)g A f g ′ . | |≤ k kpk kq Z ∗ ∗ Using the relation (Tf)g = f(T g), this is in turn equivalent to the statement T g p′ ∗ k kq ≤ A g q′ – that is the boundedness of the adjoint operator T between the dual spaces of L and Lpk respectively,k (atR least whenR 1

(Tf)g A f | |≤ k kp Z demonstrates that T may be factorised as T = Mg−1 S where S = Mg T satisfies S Lp→L1 A −1 ◦ ◦ −1k k ≤ and M −1 , the operator of multiplication by g , satisfies M −1 1 q = g ′ = 1. g k g kL →L k kq Observe that there is no obvious point of direct contact between the two regimes q 1 and 0

The result of Maurey to which we refer falls within the wider scope of Maurey–Nikisin–Stein theory, which considers factorisation of operators in a broad variety of contexts. This includes consideration of non-positive operators, sublinear operators (for example maximal functions), operators with various domains and codomains, and factorisation through various weak- and strong-type spaces, often under some auxiliary hypotheses. The particular case of positive oper- ators defined on normed lattices, taking values in Lq for q< 1, and factorising through (strong- type) L1 was considered by Maurey, however, and for this reason we refer specifically to the Maurey theory rather than the broader Maurey–Nikisin–Stein theory. For an overview of this larger theory see [26], [27], [35] and [36]. 1.2. The multilinear setting. The purpose of this paper is to develop duality and factorisation theories for certain classes of multilinear operators which are analogous to those that we have set out above in the linear setting. Amusingly, the notion of “factorisation” manifests itself in two distinct ways in our development. One of these is as a multilinear analogue of a formulation of a Maurey-type theorem as was briefly outlined in the discussion of the case 0

d d βj βj (3) (Tj Fj ) C Fj pj ≤ L (Yj ) j=1 j=1 Y Lq(X) Y where 0 <β < , 0

Strictly speaking such inequalities are multilinear only when each βj = 1; we shall nevertheless abuse language and will refer to the inequalities under consideration as “multilinear”. In fact the d case when j=1 βj = 1 will play a special role in what follows. Of course we may always assume d either thatPq = 1 or that j=1 βj = 1. To fix ideas, we discuss someP examples of inequalities falling under the scope of our study.

1 Throughout the paper, when we refer to measure spaces X, Y or Yj without explicit mention of the measure, it is implicit that the corresponding measures are µ, ν and νj respectively, unless the context demands otherwise. 2 We shall soon focus on the case pj ≥ 1 and Pj βj = 1. 4 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON

1.3. Examples. Example 1. [H¨older’s inequality] The multilinear form of H¨older’s inequality for nonnegative functions is simply

F (x) F (x)dµ(x) F p1 F pd 1 ··· d ≤k 1kL (X) ···k dkL (X) ZX d −1 where pj > 0 and j=1 pj = 1. This is of the form (3), with Tj = I for all j, q = 1 and each β = 1. But, for any fixed set of positive exponents β , it is also trivially equivalent to the j P j inequality { } β1 βd β1 βd f f q f1 fd k 1 ··· d k ≤k kq1 ···k kqd d −1 −1 for all 0 < qj < and 0

F1(π1x) Fn(πnx) dx F1 Ln−1(Rn−1) Fn Ln−1(Rn−1). b Rn ··· ≤k k ···k k Z For each 0

F1(x2)F2(x1) dx1dx2 = F1 F2. R2 R R Z Z Z In spite of its simplicity, this example will play an important guiding role for us. See Sections 6, 9.2, 9.3 and 10.2.1. The Loomis–Whitney inequality has many variants – for example Finner’s inequalities, the affine- invariant Loomis–Whitney inequality and the nonlinear Loomis–Whitney inequality. See [25], [11], and Sections 9.2 and 9.3. Example 3. [Brascamp–Lieb inequalities] The class of Brascamp–Lieb inequalities includes the n nj previous examples. Let Bj : R R be linear surjections, 1 j d. For 0

d d pj pj (4) Fj (Bj x) dx C Fj . Rn ≤ Rnj j=1 j=1 Z Y Y Z  MULTILINEAR DUALITY AND FACTORISATION 5

It is not hard to see that in order for this inequality to hold with a finite constant C, it is d necessary that j=1 pjnj = n. It is known that the constant C is finite if and only if, in addition d to j=1 pj nj =Pn, it holds that d P dim V p dimB V ≤ j j j=1 X for all V in the lattice of subspaces of Rn generated by ker B d . (See [8], [9] and [42].) { j}j=1 From this one sees easily that d kerB = 0 , d p 1, and p 1 are also necessary ∩j=1 j { } j=1 j ≥ j ≤ conditions for the finiteness of C. A celebrated theorem of Lieb [31] states that the value of the P best constant C is obtained by checking the inequality on Gaussian inputs Fj . Lieb’s theorem generalises Beckner’s theorem [6] on extremisers for Young’s convolution inequality. Suppose that 0 < r < and 0

d d pj pj (6) Fj (Bj x) dx Fj , Rn ≤ Rnj j=1 j=1 Z Y Y Z  −π|y|2 and the sharp constant 1 is achieved by the standard Gaussians Fj (y)= e . Correspondingly, in the equivalent variants presented above, the constants are also 1. The geometric Brascamp– Lieb inequalities include a suitably reformulated version of the sharp Young inequality of Beckner [6]. See Section 1.5.1 and Section 10.1 for an application of the theory we present in the context of geometric Brascamp–Lieb inequalities. Example 4. [Multilinear generalised Radon transforms] There is a vast literature on multilinear generalised Radon transforms into which we do not wish to enter. For us, this term will mean consideration of multilinear inequalities of the form (3) when the operators Tj take the form T f = f B for suitable mappings B : X Y . In most cases, X and Y will be endowed with j ◦ j j → j j 6 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON a topological or smooth structure, and the mappings Bj will respect that structure in such a way that issues of measurability do not arise. The class of multilinear generalised Radon transforms includes the Brascamp–Lieb inequalities. The most basic multilinear generalised Radon transform which is not included in the Brascamp– Lieb inequalities is probably the nonlinear Loomis–Whitney inequality. See Section 9.3 below. Example 5. [Multilinear Kakeya inequalities] The Loomis–Whitney inequality of Example 2 is equivalent to 1/(n−1) 1/(n−1) n n

aPj χPj (x) dx aPj , Rn   ≤   j=1 j=1 Z Y PXj ∈Pj Y PXj ∈Pj where is a finite family of 1-tubes in Rn which are parallel to the j’th standard basis vector Pj ej , and the aPj are arbitrary positive numbers. (A 1-tube is simply a neighbourhood of a doubly infinite line in Rn which has (n 1)-dimensional cross-sectional area equal to 1.) Multilinear − Kakeya inequalities have the same set-up, but now we allow the tubes in the family j to be approximately parallel to e , i.e. the direction e(P ) Sn−1 of the central axis of the tubePP j ∈ ∈Pj must satisfy e(P ) ej cn where cn is a small dimensional constant. Such inequalities have been studied| in [10],− [29|], ≤ [16] and [22] and have proved to be very important over the last decade with significant applications in partial differential equations and especially in number theory – see for example [12], [13], [14] and [15]. The multilinear Kakeya inequality is the statement 1/n 1/n n n

aPj χPj (x) Cn aPj .   Ln/(n−1)(Rn) ≤   j=1 j=1 Y PXj ∈Pj Y PXj ∈Pj   n   This inequality is of the form (3) with X = R , q = n/(n 1), Yj = j with counting measure, p = 1 for all j, β =1/n for all j, and T ((a ))(x)= − a χ (Px). It was Guth’s approach j j Pj Pj ∈Pj Pj Pj to such multilinear Kakeya inequalities in [29] which inspired the present paper. P The recent multilinear kj -plane Kakeya inequalities, and indeed the even more general perturbed Brascamp–Lieb inequalities, both recently established by Zhang [45], also fit into our framework, the latter as a generalisation of inequality (5).

We shall return to consider these examples in some detail later in Part III. In particular we shall discuss the affine-invariant Loomis–Whitney inequality, the nonlinear Loomis–Whitney inequal- ity and certain aspects of Brascamp–Lieb inequalities in the light of the theory we develop.

1.4. The weighted geometric mean operator. As we have just seen, all of our examples fit d pj q into the framework of inequality (3) with j=1 βj = 1, and with the L (and L ) spaces in the Banach regime, i.e. with p 1 (and q 1). We shall therefore be concerned in this paper with j P norm inequalities for the weighted≥ geometric≥ mean operator : (f ,...,f ) (T f )α1 (T f )αd Tα 1 d 7→ 1 1 ··· d d d where α = (α1,...,αd) and the αj are positive numbers satisfying j=1 αj = 1. That is, we shall consider inequalities of the form P

d d αj αj (7) (Tj fj ) A fj pj ≤ L (Yj ) j=1 j=1 Y Lq (X) Y

MULTILINEAR DUALITY AND FACTORISATION 7 for nonnegative simple functions fj (Yj ), in the regime pj 1 and q > 0. While the case q 1 is pertinent to our examples,∈ we S also wish to consider the≥ case 0

n n

βPj χPj (x) C βPj .   L1/(n−1)(Rn) ≤   j=1 Pj ∈Pj j=1 Pj ∈Pj Y X Y X The one class of examples  that does not admit a genuinely multilinear reformulation consists of the perturbed Brascamp–Lieb inequalities which were briefly mentioned in Example 5. Our first purpose in this paper is to propose and undertake a systematic study of the duality theory associated to the weighted geometric mean operator α in the context of inequality (7) in the case q 1, and some of its generalisations. It is hoped thatT the framework for this multilinear duality theory≥ will in time have applications in a wide variety of contexts. Our second purpose is to establish suitable analogues of Maurey’s theorems in the context of (7) in the case 0

1.5. A theory of multilinear duality – the regime q 1. We begin with the Banach regime q 1. ≥ ≥ One half of our duality theory – the ‘easy’ half – is largely contained in the following simple observation, the content of which is that if we have a certain pointwise factorisation property for ′ the space Lq , then the weighted geometric mean norm inequality (7) will hold.

pj q Proposition 1.1. Suppose that Tj : L (Yj ) L (X) are positive linear operators, that pj , q 1 → ′ ≥ and that d α = 1. Suppose that for every nonnegative G Lq (X) there exist nonnegative j=1 j ∈ measurable functions gj defined on X such that P d G(x) g (x)αj a.e. on X, ≤ j (8) j=1 Y ∗ and T gj ′ A G for all j. j pj q′ L (Yj ) ≤ L (X)

Then, for all nonnegative f , j ∈ Yj d d αj αj (Tj fj ) A fj ; pj ≤ L (Yj ) j=1 j=1 Lq (X) Y Y that is, (7) holds, for all nonnegative f Lp j (Y ). j ∈ j For the (easy) proof and some discussion of this result, see the more general Proposition 2.1 below. 8 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON

Rather surprisingly, the implication in Proposition 1.1 can be essentially reversed, and one of the main aims of this paper is to show that the factorisation property (8) enunciated in Proposition 1.1 is in fact necessary as well as sufficient for (7) to hold. This is the second half of the multilinear duality principle referred to in the abstract of the paper. Before coming to this, however, we note that if there is a subset of X of positive measure upon pj which Tjfj vanishes for all fj L , then this subset will play no role in the analysis of inequality (7). There is therefore no loss∈ of generality in assuming such subsets do not exist. We formalise this notion by introducing the notion of saturation below.3 In order to facilitate what follows later, we at the same time introduce the closely related notion of strong saturation, and also make the definitions in slightly greater generality than what is required by the current discussion. The definitions apply to linear operators T : (X), with a normed lattice and (X, dµ)a measure space, which are positive in the senseY→M that for every nonnY egative f we have Tf 0. (What is currently relevant is the fact that the space of simple functions defined∈ Y on a measure≥ space Y , together with the Lp norm for p 1, forms a normed lattice.) ≥ Definition 1.2. (i) We say that T saturates X if for each subset E X of positive measure, there exists a subset E′ E with µ(E′) > 0 and a nonnegative h ⊆such that Th> 0 a.e. on E′. ⊆ ∈ Y (ii) We say that T strongly saturates X if there exists a nonnegative h such that Th is a.e. bounded away from 0 on X. ∈ Y For further discussion of the relevance of these conditions, see Remarks 6 and 10 below. If T saturates a σ-finite measure space X, then there is an increasing and exhausting sequence of measurable subsets on each of which T is strongly saturating. For this and more, see Lemma 5.4 below. Now we can state one of the main results of the paper:

Theorem 1.3. Suppose that X and Yj , for j = 1,...,d, are measure spaces. Suppose that the linear operators T : (Y ) (X) are positive and that each T saturates X. Suppose that j S j → M j p 1 for all j, 1 q and d α = 1. When q = 1 suppose additionally that X is j ≥ ≤ ≤ ∞ j=1 j σ-finite. Finally, suppose that P d d αj αj (Tj fj ) A fj pj ≤ L (Yj ) j=1 j=1 Lq (X) Y Y ′ for all nonnegative simple functions f on Y , 1 j d. Then for every nonnegative G Lq (X) j j ≤ ≤ ∈ there exist nonnegative measurable functions gj on X such that d (9) G(x) g (x)αj a.e. on X, ≤ j j=1 Y and such that for each j,

(10) g (x)T f (x)dµ(x) A G q′ f j j j ≤ k kL k jkpj ZX for all simple functions fj on Yj . ∗ Remark 1. Note that we have used the formulation (10) instead of one explicitly involving Tj ∗ as we did in (8) because it is not immediately clear how Tj should be defined this context.

3For a related notion, see [44]. MULTILINEAR DUALITY AND FACTORISATION 9

The special case of Theorem 1.3 corresponding to q = 1 and G 1 can be singled out: ≡ Theorem 1.4. Suppose that X and Yj , for j = 1,...,d, are measure spaces, with X being σ- finite. Suppose that the operators T : (Y ) (X) are positive and that each T saturates X. j S j →M j Suppose that p 1, d α =1 and that j ≥ j=1 j P d d αj αj (Tj fj) dµ A fj pj X ≤ L (Yj ) Z j=1 j=1 Y Y for all nonnegative simple functions f on Y , 1 j d. Then there exist nonnegative measurable j j ≤ ≤ functions gj on X such that d 1 g (x)αj a.e. on X, ≤ j j=1 Y and such that for each j,

(11) g (x)T f (x)dµ(x) A f j j j ≤ k j kpj ZX for all simple functions fj on Yj .

In fact, Theorem 1.4 implies Theorem 1.3. Indeed, suppose that 1 < q and that ≤ ∞ d d αj αj (Tjfj ) A fj pj ≤ L (Yj ) j=1 j=1 Y Lq(X) Y

′ for all nonnegative simple functions f on Y , 1 j d. Then, for all nonnegative G Lq (X) j j ≤ ≤ ∈ with G q′ = 1, we have k kL d d αj αj (Tj fj ) G dµ A fj pj X ≤ L (Yj ) Z j=1 j=1 Y Y for all nonnegative simple functions fj on Yj , 1 j d. It is easy to see that if Tj saturates X with respect to the measure dµ, then it also does≤ so with≤ respect to G dµ. Now the measure G dµ is σ-finite irrespective of whether dµ is σ-finite measure. Therefore, by Theorem 1.4 applied with the measure G dµ in place of dµ, there are nonnegative measurable functions γj such that

d 1 γ (x)αj G dµ-a.e. on X, ≤ j j=1 Y and such that for each j,

γ (x)T f (x)G(x)dµ(x) A f j j j ≤ k jkpj ZX for all simple functions fj on Yj . Setting gj = γj G gives the desired conclusion of Theorem 1.3 when q> 1. When q = 1, factorisation of the function 1 as in Theorem 1.4 immediately yields a corresponding factorisation of each G L∞. ∈ The results described here will follow from the more general Theorem 2.2 below. 10 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON

1.5.1. An application to pointwise factorisation. As an application of Theorem 1.3, we have the following sample result concerning pointwise factorisation of nonnegative functions in L2(R2):

2 Theorem 1.5. Let v1, v2 and v3 be unit vectors in R with angle 2π/3 between each pair. Then, for every nonnegative G L2(R2), there exist nonnegative locally integrable functions g ,g and ∈ 1 2 g3 such that 1/3 1/3 1/3 G(x) g1(x) g2(x) g3(x) a.e. ≤ 2 and, for each j, for almost every line l in R which is parallel to vj ,

g dλ G j ≤k k2 Zl where dλ denotes Lebesgue measure on l. For further details, and many more results of this nature, see Section 10.1 below. 1.6. Multilinear Maurey-type factorisation – the regime 0

Theorem 1.6. Suppose that X and Yj , for j = 1,...,d, are measure spaces and that X is σ-finite. Suppose that the operators T : (Y ) (X) are positive and that each T saturates j S j →M j X. Suppose that p 1, 0

(14) g (x)T f (x)dµ(x) A f pj j j j ≤ k jkL (Yj ) ZX for all simple functions fj on Yj . We shall give the proof of this result, as a consequence of Theorem 1.4, in Section 2.3 below. Conversely, it is easy to see using H¨older’s inequality that if there exist gj such that (13) and (14) hold, then so does (12).

This result can be seen as a factorisation result in the spirit of Maurey: if we let Sj fj (x) = d d g (x)T f (x), (f ,...,f )= (S f )αj and g(x)= g (x)αj , then j j j Sα 1 d j=1 j j j=1 j = M −1 Q Tα g ◦ Sα Q where

S pj 1 A k j kL →L ≤ for all j, and −1 M −1 1 q = g ′ =1. k g kL →L k kq pj 1 In fact it has the rather strong conclusion that each Sj is bounded from L (Yj ) to L (dµ) with constant at most A (rather than the much weaker corresponding statement for the geometric mean alone). Sα MULTILINEAR DUALITY AND FACTORISATION 11

Other, different, versions of multilinear Maurey-type theorems have been studied. See for example [38] and [23]. Remark 2. One may use the classical linear Maurey–Nikisin–Stein factorisation theory of positive operators ([35], Proposition 9) to upgrade conclusions (10) of Theorem 1.3, (11) of Theorem 1.4 and (14) of Theorem 1.6. Indeed, each of these conclusions states that Tj maps into a weighted 1 pj L -space, and we can upgrade each to boundedness of Tj into a suitable weighted L -space. For example, in the context of Theorem 1.4, we may conclude that there exist nonnegative measurable functions φj on X such that

d ′ −1/(P αj /pj ) d j=1 φ (x)αj /pj dµ(x) 1  j  ≤ X j=1 Z Y   and 1/pj T f pj φ dµ A f | j j | j ≤ k j kpj ZX  for all simple fj on Yj . See also Remark 15 below. For stronger statements of this kind see the forthcoming [20]. 1.7. Structure of the paper. The paper is divided into three parts. In Part I (Sections 2–5) we present the theory of multilinear duality and factorisation and prove the main theorems. In Section 2 we state and discuss the main results at some length. The principal result is Theo- rem 2.2. Taken together with Proposition 2.1, Theorem 2.2 forms the statement of the multilinear duality principle referred to in the abstract of the paper. (Theorem 1.3 and Proposition 1.1 pre- sented in this introduction are more readily digested versions of Theorem 2.2 and Proposition 2.1 respectively.) The multilinear Maurey factorisation theorem, Theorem 2.3, is proved as a conse- quence of Theorem 2.2. (A more digestible version of Theorem 2.3 is found as Theorem 1.6 in this introduction.) In Section 3 we give a proof of a finitistic case of Theorem 2.2 which recognises and emphasises its structure as a convex optimisation or minimax problem. This perspective sets the scene for the remainder of the theoretical part of the paper. In this case, none of the functional-analytic and measure-theoretic difficulties that we encounter later are present. However, Section 3 is not strictly speaking logically necessary for the development of the theory. In Section 4 we begin to address the proof of Theorem 2.2. Our strategy will be to first consider the setting of finite measure spaces. Theorem 4.1 gives the main result in this case, and it represents a crucial step in the proof of Theorem 2.2. Already in Theorem 4.1 we are faced with substantial functional-analytic and measure-theoretic difficulties. These derive from the need to establish certain compactness statements necessary for the application of a minimax theorem. Briefly, they involve working with the of L∞(X), and dealing with various issues in the theory of finitely additive measures. In Section 5 we give the details of the proofs of Theorem 4.1 and Theorem 2.2. We begin with a couple of technical but very important lemmas. Next, we pass to the proof of the finite-measure result, Theorem 4.1, via the minimax theory. Finally, for general σ-finite X, we “glue together” factorisations obtained for subsets X of finite measure via Theorem 4.1, and we obtain the factorisations needed for Theorem 2.2. 12 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON

In a much shorter Part II, we begin to explore connections with other topics – in particular the theory of interpolation in Sections 6 and 7, and the extent to which the theory might apply in the context of more general multilinear operators in Section 8.

Finally, in Part III, we revisit the examples discussed earlier in this introduction in the light of the multilinear duality theory which has been developed. In Section 9 we give factorisation-based proofs of the affine-invariant Loomis–Whitney inequality (see Section 9.2) and the sharp nonlin- ear Loomis–Whitney inequality (see Section 9.3). In Section 10.1 we pose an interesting question related to the sharp Young convolution inequality and geometric Brascamp–Lieb inequalities, while in Section 10.2 we describe an algorithm for factorising the general Brascamp–Lieb in- equality. In Section 11 we revisit the multilinear Kakeya inequality which inspired the paper in the light of the findings of Section 8, and make an observation about the size of the constant in the finite-field version of the multilinear Kakeya inequality which is derived from our methods.

1.8. Future work. In a sequel [20] to this paper, we broaden the scope of the multilinear duality theory from positive to potentially oscillatory multilinear inequalities. In particular, we extend the multilinear duality and factorisation theorem (Theorem 2.2) to non-positive operators, and indeed refine it when the normed spaces have certain additional geometric properties (p-convexity in case of positive operators, Rademacher type in case of non-positive operators). Moreover, in forthcoming work [21], we will give an alternative proof of Theorem 2.2 which bypasses the need to consider (L∞)∗.

1.9. Acknowledgements. This paper has benefited substiantally from discussions with many individuals. In particular we should like to thank Keith Ball, Jon Bennett, Michael Christ, Michael Cowling, Alastair Gillespie, Gilles Pisier, Sandra Pott, Stuart White and Jim Wright for the insights they have shared with us on various aspects of the material of the paper. The first author is especially grateful to Michael Christ for his ongoing encouragement in the quest to find explicit factorisations. The second author is grateful to Igor Verbitsky for introducing him to the Maurey theory of factorisation in their collaboration [30] on characterising two-weight norm inequalities via factorisation. The first and third author would like to record their appreciation of the hospitality and support of the Isaac Newton Institute during the programme “Discrete Analysis” between March and July 2011, where the preliminary stages of this research were carried out and presented. The second author is supported by the Academy of Finland through funding of his postdoctoral researcher post (Funding Decision No 297929), and he is a member of the Finnish Centre of Excellence in Analysis and Dynamics Research.

Part I. Statements and proofs of the theorems

2. Statement and discussion of the main results It turns out that the theory we shall develop is most naturally presented in a more general setting. Moreover, limiting ourselves to the Lebesgue spaces Lpj and Lq in the multilinear duality theory is unnecessarily restrictive. For example, one may wish to consider multilinear inequalities of the form (7) in which the Lpj and Lq spaces are replaced by certain Lorentz spaces, Orlicz spaces or mixed-norm spaces, especially if the inequality under consideration is an endpoint inequality. We therefore introduce a more general framework in which we consider suitable spaces and d X corresponding to Lq(X) and Lpj (Y ) respectively. Thus, for α = 1, which we recall is Yj j j=1 j a standing convention, we now consider inequalities of the form P MULTILINEAR DUALITY AND FACTORISATION 13

d d αj αj (15) (Tjfj ) A fj . ≤ Yj j=1 j=1 Y X Y

Each will be an (abstract) normed lattice – such as L1 or Lpj for p 1. On the other hand, Yj j ≥ we will take to be of locally integrable4 functions defined on X, which contains the simple functions,X and is such that if f (X) and g satisfy f(x) g(x) a.e., then ∈M ∈ X | | ≤ | | f and f X g X . We may as well also assume that X is a complete measure space. Our measure∈ X spacek kX≤kwillk also be assumed to the σ-finite; these properties together then identify as a K¨othe space, see [32], Vol. II, p. 28. We shall from now on assume that is a K¨othe spaceX without further mention. Natural examples of K¨othe spaces include Lq for 1 X q . We shall need a suitable primordial dual of , denoted by ′, and defined to be ≤ ≤ ∞ X X ′ = g (X): g ′ = sup fg dµ< . X { ∈M k kX | | ∞} kfkX ≤1 Z ′ The space ′ is usually called the K¨othe dual of . If = Lq for 1 q , then ′ = Lq where 1/q +1X /q′ = 1. It is clear that ′ is a linear spaceX whichX contains≤ the≤ simple ∞ functionsX (as is contained in the class of locally integrableX functions) and is contained in the class of locally X integrable functions (as contains the simple functions). The quantity g X ′ defines a norm on ′. While by definitionX we always have the H¨older inequality k k X

fg dµ f g ′ , | |≤k kX k kX Z it may or may not be the case that ′ is norming (for ), i.e. that X X

(16) f = sup fg dµ : g ′ 1 k kX { | | k kX ≤ } Z holds for all f .5 The K¨othe dual ′ is always isometrically embedded in the norm-dual ∗, but the two spaces∈ X may not coincide inX general. X From now on, we shall adopt once and for all the convention that all named functions (f,g,h,F,G, H,β,G,S,ψ etc., often adorned with subscripts) are assumed to be nonnegative. The two excep- tions to this are the functions L and Λ appearing in the proofs of the main results.

2.1. Duality theory – easy half. The easy half of our duality theory is expressed in the following simple observation, the content of which is that if we have a certain factorisation property for the K¨othe dual ′, then the weighted geometric mean norm inequality (15) will hold. X

′ Proposition 2.1. Suppose that is a K¨othe space whose K¨othe dual is norming, and that j are normed lattices. Suppose thatXT : are positive linear operators.X Suppose furthermoreY j Yj → X

4 That is, RE |f|dµ< ∞ whenever µ(E) < ∞. 5By a result of Lorentz and Luxemburg (see [32], Vol. II, p. 29), if X is a K¨othe space, X ′ is norming if and only if X has the so-called Fatou property, that is, whenever fn ∈ X are such that fn → f a.e., with fn+1 ≥ fn ≥ 0, ∞ ′ then kfnkX → kfkX . This is automatic when X is separable. If X is L then (16) holds by inspection since X is simply L1. We shall need the notion of norming only for Proposition 2.1. 14 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON

′ that for every nonnegative G there exist nonnegative measurable functions gj on X such that ∈ X d G(x) g (x)αj a.e. on X, ≤ j j=1 (17) Y ∗ and T gj A G for all j. j ∗ ′ Yj ≤ X

Then, for all nonnegative f j ∈ Yj d d αj αj (Tj fj ) A fj . ≤ Yj j=1 j=1 Y X Y

That is, (15) holds, for all nonnegative fj j . ∈ Y ′ Proof. Take f for j =1,...,d, and G with G ′ 1. Then j ∈ Yj ∈ X k kX ≤ d d d d G(x) (T f )αj dµ(x) g (x)αj T f (x)αj dµ(x)= (g (x)T f (x))αj dµ(x) j j ≤ j j j j j j X j=1 X j=1 j=1 X j=1 Z Y Z Y Y Z Y d αj d α g (x)T f (x)dµ(x) = (T ∗g )(f ) j ≤ j j j j j j j=1 ZX  j=1 Y Y  d d d αj ∗ αj αj T gj Y∗ fj Y A G X ′ fj Y A fj ≤ k j k j k k j ≤ k k k k j ≤ k kYj j=1 j=1 j=1 Y   Y  Y where the inequalities follow in order from the first condition of (17), H¨older’s inequality, the second condition of (17), and the assumption that G X ′ 1. The proposition now follows by taking the supremum over all such G, using the factk thatk ≤′ is norming for .  X X Remark 3. If the spaces j are complete, the assumption that Tj : j is positive auto- Y 6 ∗Y → X matically implies that Tj is bounded, and so the adjoint operator Tj is well-defined. If not, we interpret the second condition of (17) as

(18) g (x)T f (x)dµ(x) A G ′ f j j j ≤ k kX k j kYj ZX for fj j and j =1,...,d, and we can still conclude the validity of (15) for functions fj j , as the∈ proof Y clearly demonstrates. ∈ Y

Remark 4. If the Tj are known to be bounded, it is immediate that (15) holds with A replaced d by T αj .7 However, the best constant A in (17) will in general be much smaller, and this j=1 k jk assertion is the main content of Proposition 2.1. Q Remark 5. Observe that Proposition 2.1 does not require any topological structure of the space X, only its nature as a measure space. Remark 6. Notice that in order for the proof to go through, we only require that the fac- torisation property – i.e. the first condition of (17) – holds for those x which contribute to d G(x) (T f )αj dµ(x) for some functions f . In other words, if a set E X with X j=1 j j j ∈ Yj ⊆ R6 Q −n n n Indeed, if not, we can find nonnegative fn with kfnk ≤ 2 but kTj fnk ≥ 2 . So for each n, 2 ≤ kTj fnk ≤ ∞ ∞ kTj (Pn fn)k ≤ C for some finite C since Pn fn ∈Yj (because Yj is a Banach space). This is a contradiction. =1 α =1 7 d j d αj This follows since k Qj=1 hj k≤ Qj=1 khj k which in turn follows from the case where each khj k = 1, which itself follows by Young’s numerical inequality and the triangle inequality. MULTILINEAR DUALITY AND FACTORISATION 15

d µ(E) > 0 has the property that for all choices f of nonnegative functions in , (T f )(x)αj = j Yj j=1 j j 0 a.e. on E, then E will play no role in the analysis. There is therefore no loss of generality in assuming such sets do not exist. So we may assume without loss of generality thatQ for all E X d ⊆ with µ(E) > 0, there exist nonnegative f such that “ T f (x)α = 0 a.e. on E” fails – j ∈ Yj j=1 j j j i.e. such that there exists E′ E, with µ(E′) > 0 such that d T f (x)α > 0 on E′. That is, ⊆ Q j=1 j j j we may assume that for all E X with µ(E) > 0, there exists E′ E with µ(E′) > 0, and, for ⊆ ′ Q ⊆ each j, a nonnegative fj j such that for all x E , Tj fj(x) > 0. This condition is equivalent to the formally slightly weaker∈ Y condition that for∈ each j, for all E X with µ(E) > 0, there exists E′ E with µ(E′) > 0 and nonnegative f such that for⊆ all x E′, T f (x) > 0.8 ⊆ j ∈ Yj ∈ j j But this is simply the statement that each Tj saturates X, as in Definition 1.2. It is unsurprising that we will require saturation when it comes to formulating and proving the converse statement. Remark 7. Note that in place of

∗ T gj A G for all j. j ∗ ′ Yj ≤ X

we could have assumed the (formally weaker) condition

d α ∗ j T gj A G . j ∗ ′ Yj ≤ X j=1 Y (A homogeneity argument shows that the two conditions are indeed equivalent.) Remark 8. Similarly, it suffices to suppose a formally weaker hypothesis (“weak factorisation”), namely that for every G ′, there exist measurable functions g on X such that ∈ X jk d G(x) g (x)αj ≤ jk j=1 Xk Y a.e. on X, and ∗ T gjk A G j ∗ ′ Yj ≤ X k X for all j. But if this holds, and if we define g j = k gjk , Minkowski’s inequality and H¨older’s inequality yield (17). So this observation does not represent a genuine broadening of the scope of Proposition 2.1. P Remark 9. The argument for Proposition 2.1 was effectively given by Guth, in a less abstract form, in his proof the endpoint multilinear Kakeya inequality [29]. However, any strategy which includes an application Proposition 2.1 to establish an inequality of the form (15) involves the potentially difficult matter of first finding a suitable factorisation. Indeed, the main work of [29] consisted precisely in finding such. In this context see also [22]. 2.2. Duality theory – difficult half. As suggested above, the implication in Proposition 2.1 can be essentially reversed, and a principal aim of this paper is to show that the factorisation property (17) enunciated in Proposition 2.1 is in fact necessary as well as sufficient for (15) to hold under very mild hypotheses. More precisely we prove: Theorem 2.2 (Multilinear duality and factorisation theorem). Suppose that (X, dµ) is a σ-finite measure space, is a K¨othe space of measurable functions on X, are normed lattices, and X Yj

8If the latter condition holds, apply it to each j in turn to obtain the former condition. 16 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON

T : (X) are positive linear maps. Suppose that each T saturates X. Suppose that j Yj →M j d d αj αj (Tj fj) A fj ≤ Yj j=1 j=1 Y X Y 9 for all nonnegative fj j , 1 j d. Then there exists a weight function w on X such that ∈ Y ′≤ ≤ 1 for every nonnegative G , there exist nonnegative measurable functions gj L (X, wdµ) such that ∈ X ∈ d (19) G(x) g (x)αj a.e. on X, ≤ j j=1 Y and such that for each j,

(20) g (x)T f (x)dµ(x) A G ′ f j j j ≤ k kX k j kYj ZX for all f . j ∈ Yj Remark 10. The hypothesis that each Tj saturates X is very natural as pointed out in Remark 6 above. Indeed, for the reasons set out there, without this hypothesis we cannot expect the conclusion to hold. Needless to say, it will play an important role in the proof of Theorem 2.2. In particular, the weight function w arises as a consequence of the saturation hypothesis. For its construction, see Section 5.3 below. If µ(X) is finite and the Tj strongly saturate X, we can take w to be the constant function 1, see Theorem 4.1 below. Remark 11. In the case d = 1 the factorisation is trivial, and (20) is simply the usual duality relation corresponding to (2).

Remark 12. If there exist gj satisfying (19) and (20), then by making one of the gj smaller if necessary, we can find gj satisfying (19) with equality in addition to (20). Remark 13. We emphasise that the constant A appearing in (20) is precisely the constant A occuring in the hypothesis. Remark 14. As in the case of Theorem 1.3, the general case of Theorem 2.2 follows from the special case in which = L1(X). Indeed, placing ourselves under the assumptions of the general case, let G ′ haveX norm 1, and observe that by H¨older’s inequality we have ∈ X d d αj αj (Tj fj) G(x)dµ(x) A fj X ≤ Yj Z j=1 j=1 Y Y for all fj j , 1 j d. This is the main hypothesis of the special case, but with respect to the measure∈ YGdµ ≤instead≤ of dµ. It is easily verified that Gdµ is a σ-finite measure, and that if Tj saturates X with respect to dµ, it also does so likewise with respect to Gdµ, and similarly for strong saturation. We may therefore conclude from the L1 case of Theorem 2.2 that there exist nonnegative measurable γj such that d γ (x)αj 1 a.e. Gdµ j ≥ j=1 Y and such that γ (x)T f (x)G(x)dµ(x) A f j j j ≤ k jkYj ZX 9i.e. a measurable function w with w(x) > 0 a.e. MULTILINEAR DUALITY AND FACTORISATION 17 for all f . Setting g = Gγ , the easy observation that j ∈ Yj j j d g (x)αj G(x) a.e. dµ j ≥ j=1 Y 1 completes the argument. (This argument does not directly place the gj in a weighted L -space, but this feature can in any case be recovered from inequality (20).) However, this observation does not simplify the proof of Theorem 2.2, and we therefore establish the general case directly.

Remark 15. Following on from Remark 2 above, if the spaces j are additionally supposed to be p -convex for some p 1, we may use the classical linear Maurey–Nikisin–SteinY theory for j j ≥ positive operators to upgrade conclusion (20) of Theorem 2.2 to boundedness of each Tj into a suitably weighted Lpj -space. A similar remark applies in the context of Theorem 1.6 below. This perspective is further explored in [20].

The proof of Theorem 2.2 is highly nonconstructive and comes about as a result of duality methods in the theory of convex optimisation which ultimately rely upon a form of the minimax principle. For the details of the proof see Sections 3, 4 and 5 below. Nevertheless, in some cases, constructive factorisations can be given, and in other cases, the existence of the factorisation raises interesting questions and links with other areas of analysis. See Sections 6, 8, 9.2, 9.3 and 10.

2.3. Multilinear Maurey-type theory. In this section we state and prove a slight general- isation of Theorem 1.6, using the case = L1(X) of Theorem 2.2. Interestingly, the classical Maurey theorem follows easily from TheoremX 2.2 specialised to the bilinear case d = 2 in which one of the normed lattices is one-dimensional. (Therefore, the case d = 1 of what follows is not trivial, in contrast to the situation for Theorem 2.2.) Theorem 2.3 (Multilinear Maurey-type theorem). Suppose (X, dµ) is a σ-finite measure space, j are normed lattices, and Tj : j (X) are positive linear maps. Suppose that each Tj Ysaturates X. Let 0

d d αj αj (21) (Tjfj ) A fj ≤ Yj j=1 j=1 Y Lq(X) Y

for all nonnegative fj j , 1 j d. Then there exist nonnegative measurable functions gj on X such that ∈ Y ≤ ≤ d αj (22) g (x) q′ =1 k j kL (X) j=1 Y and such that for each j,

(23) g (x)T f (x)dµ(x) A f j j j ≤ k jkYj ZX for all f . j ∈ Yj It is an easy exercise using H¨older’s inequality to show that if there exist gj such that (22) and (23) hold, then (21) also holds. As in the case of Theorem 1.6, Theorem 2.3 admits an interpretation as a statement about factorisation of operators, see Section 1.6. 18 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON

Proof. The main hypothesis is that

d d αj q αj q q (Tj fj ) dµ A fj X ≤ Yj Z j=1 j=1 Y Y for all f . Let β = α q for 1 j d and let β = 1 d β = 1 q > 0. Let j ∈ Yj j j ≤ ≤ d+1 − j=1 j − Y = 0 and let be the trivial normed lattice R defined on the singleton measure space d+1 { } Yd+1 P 0 . Let Td+1 : d+1 (X) be the λ λ1 where 1 denotes the constant function {taking} the valueY 1 on→MX. Then we have 7→

d+1 d+1 βj βj q (Tj fj) dµ A fj X ≤ Yj Z j=1 j=1 Y Y for all f . So by Theorem 2.2 in the case = L1(X) (and with d + 1 in place of d, see also j ∈ Yj X Remark 12 above), we conclude that there exist measurable functions G1,...,Gd+1 such that

d+1 βj (24) Gj (x) = 1 a.e. on X, j=1 Y and such that for each 1 j d + 1, ≤ ≤ (25) G (x)T f (x)dµ(x) Aq f j j j ≤ k jkYj ZX for all f . j ∈ Yj 1−q For 1 j d, set gj (x)= A Gj (x); then (25) immediately gives (23) for 1 j d. By (24) we have≤ ≤ ≤ ≤ d αj 1−q (q−1)/q gj(x) = A Gd+1(x) a.e. on X, j=1 Y while (25) for j = d + 1 gives

G (x)dµ(x) Aq. d+1 ≤ ZX Combining these last two relations gives (22) as desired. 

3. Discrete case: a convex optimisation problem 3.1. Basic set-up. The idea behind the proof of Theorem 2.2 is to view problem (19) and (20) as a convex optimisation problem. That is, we replace the number A in (20) by a variable K and seek to minimise over K. To illustrate how this works, we first prove the theorem in a model case 1 when X and Yj are finite sets endowed with counting measure and j = L (Yj ) for j =1,...,d. One reason for doing this case first is that there are no measure-theoreticY or functional-analytical difficulties to be dealt with in this setting, and indeed ′ = ∗ is simply the class of all functions defined on X with the norm dual to that of . It thereforeX X allows us to emphasise the nature of the problem as one concerning convex optimisation.X MULTILINEAR DUALITY AND FACTORISATION 19

The minimisation problem we propose to examine now reads as follows. Fix G : X [0, ) and consider → ∞ γ = inf K K,gj d αj (26) such that G(x) gj(x) for all x X and ≤ ∈ j=1 ∗Y max Tj gj (yj ) K G X ∗ for all j =1,...,d. yj ∈Yj ≤ k k We note that this is a convex optimisation problem since we are minimising a convex, in fact linear, function on the convex domain consisting of the (d + 1)-tuples (K,gj) satisfying the constraints in (26). The convexity of this domain follows from the fact that the second set of inequalities is linear in the arguments K and gj , and the operation of taking the geometric mean on the right hand side of the first set of inequalities is a concave function. We note that the set of (K,gj) satisfying the constraints in (26) is not empty and that we can in fact find (K,gj) satisfying these constraints with strict inequality by taking each gj to be 2G + 1 and letting K be sufficiently large. Thus problem (26) satisfies what is known as Slater’s condition. (We do not give full details here as the discussion will eventually be subsumed into that of the next section.) In particular we certainly have γ < + . ∞ We therefore follow a standard approach to convex optimisation problems, see for example [17]. We introduce Lagrange multipliers ψ and h , where ψ : X R (for the first set of constraints), j → + and hj : Yj R+ (for the second set). Note that we are only interested case where these functions take→ nonnegative values since each of the constraints is an inequality constraint. We then introduce the Lagrangian functional

d d αj ∗ (27) L = K + ψ(x) G(x) g (x) + h (y )(T g (y ) K G ∗ ).  − j  j j j j j − k kX j=1 j=1 xX∈X Y X yXj ∈Yj We emphasise that this function and the corresponding one defined in the proof of the general case are the only functions which we allow to take negative values. 10 For nonnegative K, gj, ψ, and hj we now consider the two problems

γL = inf sup L and η = sup inf L K,g K,g j ψ,hj ψ,hj j called the primal problem and the dual problem respectively. We shall show that (i) the problem for γ is identical to the problem for γ , (ii) η A where A is any number such that inequality L ≤ (15) holds, and (iii) η = γL (it is obvious that η γL). Finally, we show that the infimum in the definition of γ is attained, and this will complete≤ the proof of the theorem in the special case. 3.2. Identification of the problems for γ and γ . We begin by studying γ . Fix K 0 and L L ≥ gj and consider supψ,hj L. Suppose that any of the conditions in (26) is not satisfied at some point. Then take the relevant function ψ or hj for some j to have value t> 0 at a point where an inequality fails and let all of the functions be zero everywhere else. Then let t and notice that sup L goes to + since t is multiplied by a positive number. So if sup→ ∞L< + we ψ,hj ∞ ψ,hj ∞ must have that the conditions of (26) are satisfied. Conversely, if these conditions are satisfied then all factors multiplying ψ(x) and hj(yj ) for any x and yj are non-positive so the supremum is attained by taking them all to equal 0. So, for each fixed (K,gj ), we have that sup L< + ψ,hj ∞ if and only if the conditions in (26) hold, in which case, supψ,hj L = K. Thus we see that the

10The subscript L in γL indicates that we are looking at the Lagrangian version of the problem as opposed to the original version which has γ without a subscript. 20 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON problem for γL is identical to problem (26), yielding γL = γ. Moreover the infimum in the definition of γ is attained if and only if the infimum in the definition of γL is attained.

3.3. Proof that η A. We rearrange L as follows: ≤ d L = ψ(x)G(x)+ K 1 G ∗ h (y )  −k kX j j  x∈X j=1 yj ∈Yj (28) X X X d  d  + g (x)T h (x) g (x)αj ψ(x)  j j j − j  j=1 j=1 xX∈X X Y   Let us fix ψ and h and consider inf L. First of all, note that inf L = unless j K,gj K,gj −∞ d (29) G ∗ h (y ) 1 k kX j j ≤ j=1 X yXj ∈Yj since if this inequality fails then the term multiplying K in L is negative and so by taking gj =0 and letting K go to infinity we get that inf L = . Also note that inf L = unless K,gj −∞ K,gj −∞ d (30) ψ(x) (α−1T h (x))αj ≤ j j j j=1 Y for all x X. Seeing this is a matter of choosing g (x) to balance the arithmetic-geometric mean ∈ j inequality. Specifically, suppose that this condition (30) fails at a point x0. Then we let K = 0 and gj(x) = 0 for all x = x0 and all j = 1,...,d. There are now two cases to consider. Firstly, if there exists an index 6j such that T h (x ) = 0 then we take g (x ) = 1 for all j = j and 0 j0 j0 0 j 0 6 0 gj0 (x0)= t> 1. Then

d α L = ψ(x)G(x)+ T h (x ) t j0 ψ(x ). j j 0 − 0 x∈X j=1 X jX6=j0 Since α > 0 we can let t go to infinity and see that inf L = . In the other case we have j0 K,gj −∞ that Tjhj (x0) > 0 for all j =1,...,d. Then we let

d −1 −1 αj′ gj (x0)= tαj ((Tj hj )(x0)) (αj′ Tj′ hj′ )(x0) ′ jY=1 and note that d L = ψ(x)G(x)+ t (α−1T h (x ))αj ψ(x ) .  j j j 0 − 0  j=1 xX∈X Y So by the assumption of the failure of (30) atx we see that letting t yields inf L = . 0 → ∞ K,gj −∞ Conversely, if conditions (29) and (30) hold then the factor multiplying K is nonnegative and an application of the arithmetic-geometric mean inequality gives that for any choice of gj then for each x X the term in the second bracket of (28) is nonnegative, so we attain infK,gj L by letting K = 0 and∈ g (x) = 0 for all x X and j =1,...,d. Hence, for each fixed (ψ,h ), inf L> j ∈ j K,gj −∞ if and only if ψ and hj satisfy conditions (29) and (30), in which case infK,gj L = x∈X ψ(x)G(x). P MULTILINEAR DUALITY AND FACTORISATION 21

Noting that there always exist ψ and hj satisfying conditions (29) and (30), we see that η is the solution to η = sup ψ(x)G(x) ψ,hj xX∈X d such that ψ(x) (α−1T h (x))αj for all x X and (31) ≤ j j j ∈ j=1 Y d G ∗ h (y ) 1. k kX j j ≤ j=1 X yXj ∈Yj For any ψ and hj satisfying the conditions in (31) we can calculate

d d −1 αj −1 αj ψ(x)G(x) (αj Tjhj (x)) G(x) (αj Tjhj ) G ≤ ≤ X ∗ x∈X x∈X j=1 j=1 X X X Y Y d d −1 αj A α hj G X ∗ A hj 1 G X ∗ A ≤ k j kY| k k ≤ k k k k ≤ j=1 j=1 Y X where the inequalities follow in order from the first condition of (31), the definition of the norm on ∗, the inequality (15), the arithmetic-geometric mean inequality, and the second condition of (X31). Taking the supremum now yields η A. ≤

3.4. Proof that γL = η and existence of minimisers. This is a minimax argument. As we have noted above, it is immediate that η γ and this is referred to as weak duality. The other ≤ L direction, giving γL = η, is called strong duality and does not hold in general. However there are various conditions which guarantee strong duality, such as Slater’s condition which is the condition that the original problem (26) is convex and there exists a point satisfying all of the constraints with strict inequality. See [17], p.226. We have noted above that Slater’s condition holds in our setting. Moreover, Slater’s condition guarantees the existence of a maximiser for the dual problem. However, we need optimisers for the primal problem. If for all x X we have ∈ Tj 1(x) > 0 – which is simply the saturation hypothesis in our present case – then the set of gj ’s which satisfy the constraints of (26) with K = 2A will be compact, and therefore a minimiser will exist.

4. General case: overview of the proof Let us now turn to the argument for Theorem 2.2 in the general case. It will entail substantial measure-theoretic and functional-analytic considerations not present in the case when X and Yj are finite sets. While it is an attractive idea to try to establish Theorem 2.2 by approximating the general case by the discrete case, this does not seem a feasible route, even when and X Yj are Lq and Lpj spaces respectively, and a direct approach is therefore required. The bulk of the proof of Theorem 2.2 will be devoted to establishing a special case in which X is a finite measure 11 space and where we impose strong saturation on the Tj instead of saturation. This leads to the 1 crucial conclusion that we can take the factors gj to lie in L (X, dµ). The result reads as follows: Theorem 4.1. Suppose X is a finite measure space, is a K¨othe space of functions defined on X, are normed lattices, and that the linear operatorsX T : (X) are positive. Suppose Yj j Yj →M

11To be clear, a measure space (X, dµ) with µ(X) < ∞, not a finite set X with counting measure. 22 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON that each Tj strongly saturates X. Suppose that

d d αj αj (32) (Tj fj) A fj ≤ Yj j=1 j=1 X Y Y 12 ′ holds for all nonnegative fj j , 1 j d. Then for every nonnegative G there exist nonnegative functions g L∈1(X, Y dµ)≤such≤ that ∈ X j ∈ d (33) G(x) g (x)αj a.e. on X, ≤ j j=1 Y and such that for each j,

(34) g (x)T f (x)dµ(x) A G ′ f j j j ≤ k kX k j kYj ZX for all f . j ∈ Yj In the proof of this theorem we introduce an extended real-valued Lagrangian function L(Φ, Ψ) (where Φ corresponds to the variables (K,gj) and Ψ corresponds to the variables (ψ,hj ) of the discrete model case discussed above). See Section 5.2 below for precise details of the definition of L. As in the model case, we relate supΨ infΦ L (which we had previously called η) to problem (33) and (34) and infΦ supΨ L (which we had previously called γ) to inequality (32). We then need to show that min sup L(Φ, Ψ) = sup inf L(Φ, Ψ), Φ Ψ Ψ Φ and for this we need to use the Lopsided Minimax Theorem (which can be found as Theorem 7 from Chapter 6.2 of [2]): Theorem 4.2. Suppose C and D are convex subsets of vector spaces and that C is endowed with a topology for which the vector space operations are continuous. Further, suppose that L : C D R satisfies × → (i) Φ L(Φ, Ψ) is convex for all Ψ D; (ii) Ψ 7→ L(Φ, Ψ) is concave for all Φ∈ C; (iii) Φ 7→ L(Φ, Ψ) is lower semicontinuous∈ for all Ψ D; and 7→ ∈ (iv) there exists a Ψ0 D such that the sublevel sets Φ C : L(Φ, Ψ0) λ are compact for all sufficiently∈ large λ R. { ∈ ≤ } ∈ Then (35) min sup L(Φ, Ψ) = sup inf L(Φ, Ψ). Φ∈C Ψ∈D Ψ∈D Φ∈C Remark 16. The existence of the minimum on the left-hand side of (35) is part of the conclusion: there exists a Φ¯ C such that sup L(Φ¯, Ψ) = inf sup L(Φ, Ψ). Once we know that ∈ Ψ∈D Φ∈C Ψ∈D inf sup = sup inf this is easy because (iii) tells us that the map Φ supΨ∈D L(Φ, Ψ) is lower semicontinuous, and (iv) then tells us that the sublevel sets 7→

Φ C : sup L(Φ, Ψ) λ Φ C : L(Φ, Ψ0) λ { ∈ Ψ∈D ≤ }⊆{ ∈ ≤ } are closed and compact, and hence Φ supΨ∈D L(Φ, Ψ) achieves its minimum on any such set. The fact that sup inf inf sup is trivial,7→ so the main content of the theorem is that inf sup sup inf. ≤ ≤

12Notice that a hypothesis of strong saturation is unrealistic in the presence of inequality (32) unless X has finite measure. MULTILINEAR DUALITY AND FACTORISATION 23

Remark 17. There is nothing to stop both sides of (35) from being + . Indeed, a nontrivial conclusion of the theorem is that if the right-hand side is finite, so is the∞ left-hand side. Remark 18. Traditional versions of minimax theorems assume that C itself is compact, rather than compactness of certain sublevel sets as condition (iv). However, in our case, we cannot, for the reasons set out below, expect C to be compact. It is a remarkable feature of our analysis that the saturation hypothesis we must impose corresponds precisely to condition (iv) of the minimax theorem. Remark 19. The observant reader will have noticed that we have indicated our intention to introduce an extended real-valued Lagrangian L, but the minimax theorem applies only to real- valued Lagrangians. This mismatch necessitates a small detour which we wish to suppress here.13 For details see Section 5.2.1 below.

It is a somewhat delicate matter to choose the vector space where we will locate the variables Φ featuring in the Lagrangian which we will use. Corresponding to the variables gj occuring in the discrete model case of Section 3, we will now have variables Sj , which we would like to take to be 1 elements of L (X)+. (It is the weighted geometric mean of a particular collection of these which will ultimately furnish the desired factorisation.) However, it turns out to be helpful to instead ∞ ∗ allow, in the first instance, the Sj be elements of the larger space L (X)+, that is, the positive cone of the dual of L∞(X).14 Thus we consider the vector space R L∞(X)∗ L∞(X)∗ and take C to be a suitable subset of the positive cone in this space.× Ideally we×···× would like to take C to be a norm-bounded convex subset and then use the Banach–Alaoglu theorem to assert compactness of C; but since we are not expecting any quantitative L1 bounds on the functions Sj appearing in the factorisation, there is no natural norm-bounded set with which to work. R ∞ ∗ ∞ ∗ Instead, we take C to be the whole positive cone + L (X)+ L (X)+, endowed with the weak-star topology. The price for this is the need× to verify hyp×···×othesis (iv) of Theorem 4.2. Fortunately this turns out to be not so difficult, and in fact is rather natural in our setting. ∞ ∗ Carrying out this process will yield some distinguished members of L (X)+. However, working ∞ ∞ ∗ with the dual of L (X) presents its own difficulties since some elements of L (X)+ are quite exotic. Fortunately the theory of finitely additive measures comes to the rescue, and we will be able to show that elements satisfying the properties we require can be in fact be found in the 1 smaller space L (X)+. See Section 5.2 below for more details. To set the scene for this, we recall three results of Yosida and Hewitt which can be found in [43]. The setting for each of these results is a σ-finite measure space (X, dµ). Theorem 4.3. There is an isometric isomorphism between the space of finitely additive measures on X of finite total variation which are µ-absolutely continuous15 and the space of bounded linear functionals on L∞(X, dµ). For a finitely additive measure τ with these properties the ∞ ∗ corresponding element of L (X, dµ) is given by τ(ψ) = X ψ dτ (where the integral is the so- called Radon integral). Furthermore L1(X, dµ) embeds isometrically into L∞(X, dµ)∗ in such a 1 R ∞ way that the application of g L (X, dµ) to an element of ψ L (X, dµ) is given by X ψg dµ where the integral is now the∈ Lebesgue integral. ∈ R 13B. Ricceri has recently informed us (private communication) that Theorem 4.2 continues to hold when the Lagrangian is permitted to take the value +∞. The detour takes no longer than establishing this more general minimax statement. 14 1 Had we instead opted to work from the outset with Sj ∈ L (X)+, we would have been forced to place unnatural 1 topological conditions on X in order to identify L (X)+ with a subspace of a dual space, and in any case we would have to work in the larger space of finite regular Borel measures on X in order to exploit weak-star compactness. 15This means that the finitely additive measure τ satisfies τ(E) = 0 whenever µ(E)= 0. 24 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON

∞ ∗ Theorem 4.4. Any element S L (X, dµ) can be written uniquely as S = Sca + Spfa where ∈ 1 Sca is countably additive (and hence is given by integration against a function in L (X, dµ)) and S is purely finitely additive. Furthermore S 0 if and only if S 0 and S 0. pfa ≥ ca ≥ pfa ≥

We need not concern ourselves here with the definition of purely finitely additive measures, but, in order to be able to use these results, we do need a useful characterisation of which measures are purely finitely additive.

Theorem 4.5. A nonnegative finitely additive measure τ which is µ-absolutely continuous is purely finitely additive if and only if for every nonnegative countably additive measure σ which is µ-absolutely continuous, every measurable set E and every pair of positive numbers δ1 and δ2, there is a measurable subset E′ of E such that σ(E′) <δ and τ(E E′) <δ . 1 \ 2

We wish to remark that analysis related to the dual space of L∞ has also been employed in a number of other contexts recently. See for example [3], [39], [41] and, in the financial mathematics literature, [34].

5. General case: details of the proof 5.1. Preliminaries. We shall first need two lemmas which will be useful for the proof of Theo- rem 4.1 and also that of Theorem 2.2 itself. The first one is they key technical tool which, in the context of Theorem 4.1, will allow us do induce existence of suitable integrable functions from ∞ existence of corresponding members of the dual of L . We shall continue to assume that αj > 0 d and that j=1 αj = 1. P ∞ ∗ Lemma 5.1. Let (X, dµ) be a σ-finite measure space and suppose that Sj L (X)+. Suppose that G is a measurable function such that ∈

d d (36) G(x) βαj (x) dµ(x) α S (β ) for all simple functions β on X. j ≤ j j j j X j=1 j=1 Z Y X

If (Sjrn) denotes the Radon–Nikodym derivative with respect to µ of the component of Sj which is countably additive, then

d (37) G(x) S (x)αj a.e. on X. ≤ jrn j=1 Y Conversely, if G is such that (37) holds, then (36) holds.

Remark 20. This result extends the special case d = 1 which is implicit in Theorem 4.4.

Proof. The converse statement follows immediately from the arithmetic-geometric mean inequal- ity, so we turn to the forward assertion. Suppose not. Then there exists a set E0 with µ(E0) > 0 such that inequality (37) fails on E0 and we can find an ε> 0 and E1 E0 with µ(E1) > 0 such that ⊆ d G(x) ε> Sαj (x) − jrn j=1 Y MULTILINEAR DUALITY AND FACTORISATION 25 for all x E . Now take β to be simple functions supported on E . Then we get ∈ 1 j 1 d d d αj αj αj Sjrn(x) βj (x) dµ(x)+ ε βj (x) dµ(x) E1 j=1 j=1 E1 j=1 Z Y Y Z Y d d α (38) G(x) β j (x) dµ(x) α S (β ) ≤ j ≤ j j j E1 j=1 j=1 Z Y X d d = αj Sjrn(x)βj (x) dµ(x)+ αj βj (x) dτjpfa j=1 E1 j=1 E1 X Z X Z where τjpfa is the purely finitely additive measure associated to the purely finitely additive component Sjpfa of Sj .

We can find a subset E2 E1 with µ(E2) > 0 and a C > 0 such that Sjrn(x) C for all x E2 and all j. There are now⊆ two cases to consider. ≤ ∈

First, assume that there exists a subset E E such that µ(E ) > 0 and an index j such that 3 ⊆ 2 3 0 Sj0rn(x) = 0 for all x E3. Now let δ be small and positive (to be specified later) and E4 a ∈ 16 subset of E3 with 0 < µ(E4) < (also to be specified later ), and take βj = δχE4 for j = j0 −1 ∞ 6 1−α d αj and β = δ j0 χ . This implies that β (x) = 1 for all x E and so the top line j0 E4 j=1 j ∈ 4 of (38) equals 0 + εµ(E ). The first term on the bottom line of (38) can be bounded by Cδµ(E ) 4 Q 4 since there is no contribution from the term with index j0. The second term we can bound by 1−α−1 j0 δ j τjpfa (E4). We have not chosen E4 precisely yet. To do this we use Theorem 4.5 above. P 

Indeed, applying Theorem 4.5 with τ := j τjpfa, σ := µ, E := E3 and δ1 := µ(E3)/2 gives that for all δ > 0 there is an E E such that µ(E E ) < µ(E )/2 and τ(E ) < δ . So (38) 2 4 3 P 3 4 3 4 2 implies ⊆ \ 1−α−1 1−α−1 εµ(E ) Cδµ(E )+ δ j0 τ(E ) Cδµ(E )+ δ j0 δ . 4 ≤ 4 4 ≤ 4 2 1−α−1 Now choose δ = εµ(E )/4δ j0 , so that for some E E we have 2 3 4 ⊆ 3 εµ(E ) Cδµ(E )+ εµ(E )/4 Cδµ(E )+ εµ(E )/2. 4 ≤ 4 3 ≤ 4 4 Finally, choosing δ <ε/(2C) yields a contradiction, since by construction µ(E4) > 0.

In the other case, we have that Sjrn(x) > 0 for a.e. x E2 and all j. Then we can find a subset E E with 0 < µ(E ) < and a number c> 0 such∈ that S (x) c for all x E and all 3 ⊆ 2 3 ∞ jrn ≥ ∈ 3 j. We define uj on this set as

d −1 αk uj(x)= Sjrn(x) Skrn(x) kY=1 and note that u (x) c−1C and that uαj (x) = 1. Since these functions are bounded then if j ≤ j j we are given δ > 0 we can find simple functions β˜ such that u (x) δ β˜ (x) u (x) for all Q j j − ≤ j ≤ j x E . We may assume that β˜ (x) cC−1 for all x E . Let us take β = β˜ χ where E is ∈ 3 j ≥ ∈ 3 j j E4 4

16 We are using σ-finiteness of µ to ensure that we can find such an E4 with µ(E4) finite. 26 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON a subset of E to be chosen. Then for x E we have that 3 ∈ 4 d d d k d 0 uαj (x) βαj (x)= uαj (x) (uαk (x) βαk (x)) βαj (x) ≤ j − j  j k − k j  j=1 j=1 j=1 Y Y Xk=1 Y j=Yk+1 d k  d  ξαk (x) = uαj (x)α k (u (x) β (x)) βαj (x) dC2c−2δ.  j k ξ (x) k − k j  ≤ j=1 k Xk=1 Y j=Yk+1   Here ξj0 (x) lies between βj0 (x) and uj0 (x) and we have used that c/C βj (x), ξj (x),uj (x) C/c. Now we can estimate the first term on the top line of (38) from below≤ by ≤

d d d Sαj (x) uαj (x) dC2c−2δ dµ(x) Sαj (x) dµ(x) µ(E )dC3c−2δ, jrn  j −  ≥ jrn − 4 E4 j=1 j=1 E4 j=1 Z Y Y Z Y   2 −2 and the second term on the top line of (38) we can estimate from below by εµ(E4)(1 dC c δ) αj − since j uj (x)=1on E4. The first term on the bottom line of (38) we can estimate from above using β u on E and Q j ≤ j 4 the definition of uj by d αj Sjrn(x) dµ(x). E4 j=1 Z Y −1 The second term we can bound by Cc j τjfa (E4). Collecting this we have that P  d Sαj (x) dµ(x) µ(E )dC3c−2δ + εµ(E )(1 dC2c−2δ) jrn − 4 4 − E4 j=1 Z Y d Sαj (x) dµ(x)+ Cc−1 τ (E ). ≤ jrn  jpfa 4 E4 j=1 j Z Y X The integrals cancel17 and we get   εµ(E ) µ(E )(εdC2c−2δ + dC3c−2δ)+ Cc−1τ(E ) 4 ≤ 4 4 where τ = j τjpfa. We can then choose E4 and δ in much the same way as before to yield a contradiction. Note that δ will only depend on d, C, c and ε. P So in both cases we have a contradiction to the existence of E0, and so (37) must hold. 

∞ ∗ ∞ It turns out that we shall need to consider the action of S L (X, dµ)+ not just on L (X), but on general nonnegative measurable functions in (X).∈ The reasons for this are explained in Section 5.2 below. To this end, we extend S to (MX) by declaring, for F (X) , M + ∈M + S(F ) := sup S(f):0 f F,f L∞(X) = sup S(φ):0 φ F, φ simple . { ≤ ≤ ∈ } { ≤ ≤ } Of course S(F ) will often now take the value + . ∞ The second lemma concerns continuity properties of this extension. Consider the map S S(F ) ∞ ∗ ∞ 7→ for fixed F (X)+ as S ranges over L (X)+. If F L (X) this map is norm continuous and hence weak-star∈ M continuous. For F (X) we can∈ assert less. ∈M + 17 ∞ ∗ 1 Since Sj ∈ (L ) we have Sjrn ∈ L , and the terms we are cancelling are indeed finite. MULTILINEAR DUALITY AND FACTORISATION 27

∞ ∗ R Lemma 5.2. Fix F (X)+. Then the map S S(F ) from L (X)+ to + is weak-star lower semicontinuous.∈ M 7→ ∪ { ∞}

∞ ∗ We remark that we have to be cautious here since L (X)+ with the weak-star topology is not a metric space; so we cannot simply concern ourselves with sequential lower semicontinuity.

∞ ∗ Proof. Let S L (X)+. Either S(F )=+ or S(F ) < + . Let us first deal with the latter case. We need∈ to show that for every ǫ> 0 there∞ is a weak-star∞ open neighbourhood U of S such that for R U we have R(F ) S(F ) ǫ. ∈ ≥ − Since S(F ) < + there is an f L∞(X) with 0 f F such that S(f) >S(F ) ǫ. Let ∞ ∈ ≤ ≤ − U = R L∞(X)∗ : R(f) >S(F ) ǫ . { ∈ + − } Then S U, and U is weak-star open since for each f L∞(X) the functional R R(f) is weak-star∈ continuous. So for R U we have ∈ 7→ ∈ R(F ) R(f) >S(F ) ǫ ≥ − which is what we needed. Now we look at the case S(F )=+ . We now need to show that for every N N there is a weak-star open neighbourhood U of ∞S such that for R U we have R(F ) N. ∈ ∈ ≥ Since S(F )=+ there is an f L∞(X) with 0 f F such that S(f) >N. Let ∞ ∈ ≤ ≤ U = R L∞(X)∗ : R(f) >N . { ∈ + } Then S U, and U is weak-star open since for each f L∞(X) the functional R R(f) is weak-star∈ continuous. So for R U we have ∈ 7→ ∈ R(F ) R(f) >N ≥ which is what we needed.  5.2. Proof of Theorem 4.1. Suppose we are in the situation in the statement of Theorem 4.1. ′ In particular, we assume that G , and we may clearly assume that G ′ = 0. ∈ X k kX 6 We recall from Section 4 that we take C (in which we locate the variables Φ = (K,Sj )) to R ∞ ∗ d R ∞ ∗ d be the positive cone + (L (X)+) in the vector space (L (X) ) , and C is given the topology inherited from the× product topology of the correspondin× g weak-star topologies. We take D (in which we locate the variables Ψ = (βj ,hj )) to be the positive cone in the vector space (X)d . S × Y1 ×···×Yd R ∞ ∗ Therefore, for K +, Sj L (X)+, βj simple functions on X and hj j we consider the functional ∈ ∈ ∈ Y d d L = K + G(x) βαj (x) dµ(x) α S (β )  j − j j j  X j=1 j=1 Z Y X d  + S (T h ) K G ′ h . j j j − k kX k jkYj j=1 X  Note that the integral term is well-defined since G ′ (which is contained in L1(X, dµ) when ∈ X µ is a finite measure, as we have previously observed) and the βj are simple functions, and that the terms Sj (βj ) are also well-defined since the βj are bounded functions. The terms Sj (Tj hj) are well-defined via the extension of S to (X) as discussed in Section 5.1 above. Thus j M + 28 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON

L : C D R + is well-defined and takes values in R + , with the possible value + arising× → when∪{T h∞}is not a bounded measurable function. ∪{ ∞} ∞ j j We next want to see how we can apply the minimax theorem, Theorem 4.2 to this Lagrangian. Recall from Remark 19 above that we have a problem in so doing, since our Lagrangian may take the value + , while Theorem 4.2 requires that the Lagrangian be real-valued. To be clear, what we desire –∞ and what we shall indeed obtain – is the conclusion min sup L(Φ, Ψ) = sup inf L(Φ, Ψ) Φ∈C Ψ∈D Ψ∈D Φ∈C of Theorem 4.2 in our case, but in order to achieve this we need to make a detour. 5.2.1. A detour. We now describe the necessary detour. This involves modifying the Lagrangian we have defined in order to make it real-valued, but without altering its essential purpose. The main technical difference is that instead of allowing Sj to act on the possibly unbounded Tjhj , we have it act on an arbitrary nonnegative simple function ψ satisfying ψ T h , j j ≤ j j We therefore introduce a new Lagrangian Λ : C D˜ R, where C is as before, and where × → D˜ = (β ,h , ψ ) (X)d (X)d : β 0, h 0, 0 ψ T h . { j j j ∈ S × Y1 ×···×Yd × S j ≥ j ≥ ≤ j ≤ j j } Note that D˜ is convex. For (K,S ) C and (β ,h , ψ ) D˜ we define j ∈ j j j ∈ d d Λ= K + G(x) βαj (x) dµ(x) α S (β )  j − j j j  X j=1 j=1 Z Y X d  + S ψ K G ′ h . j j − k kX k jkYj j=1 X  Note that Λ is real-valued since Sj ψj is real-valued. Moreover, note that by the definition of the extension of S to (X) , we have j M + (39) L((K,Sj), (βj ,hj )) = sup Λ((K,Sj ), (βj ,hj , ψj )). {ψj : ψj ≤Tj hj } We will momentarily check that the Lagrangian Λ satisfies the hypotheses of Theorem 4.2, but taking this as read for now, we deduce using (39) that min sup L = min sup Λ = sup inf Λ. (K,Sj )∈C (K,Sj )∈C (K,Sj )∈C (βj ,hj )∈D (βj ,hj ,ψj )∈D˜ (βj ,hj ,ψj )∈D˜

But since trivially sup inf inf sup, we have, using (39) once more, ≤ sup inf Λ sup inf sup Λ = sup inf L. (K,Sj )∈C ≤ (K,Sj )∈C (K,Sj )∈C (βj ,hj ,ψj )∈D˜ (βj ,hj )∈D ψj ≤Tj hj (βj ,hj )∈D Combining the last two displays we obtain min sup L sup inf L. (K,Sj )∈C (βj ,hj )∈D ≤ (βj ,hj )∈D (K,Sj )∈C Since the reverse inequality is once again trivial we conclude that min sup L(Φ, Ψ) = sup inf L(Φ, Ψ) Φ∈C Ψ∈D Ψ∈D Φ∈C as we needed. Now we need to look at conditions (i) – (iv) of Theorem 4.2 in our case where Λ replaces L. Concerning (i), the map S S(F ) is linear on L∞(X)∗ for each fixed F (X). Therefore, 7→ ∈ S MULTILINEAR DUALITY AND FACTORISATION 29 for each fixed Ψ˜ D˜, the map Φ Λ(Φ, Ψ)˜ is affine, thus convex on C. Concerning (ii), the ∈ 7→ ∞ ∗ map F S(F ) is linear and hence concave on (X) for each fixed S L (X)+. Moreover the geometric7→ mean is a concave operation and theS map h h is convex.∈ Therefore, for each 7→ k kYj fixed Φ C, the map Ψ˜ Λ(Φ, Ψ)˜ is concave on D˜. Concerning (iii), this follows directly from the norm-continuity∈ of S7→ S(F ) on L∞(X)∗ for each fixed F (X). 7→ ∈ S Condition (iv) is more interesting, and it is in verification of this condition that we use the crucial strong saturation hypothesis of Theorem 4.1. We need to see that for some Ψ˜ D˜ the sublevel 0 ∈ sets Φ C : L(Φ, Ψ˜ ) λ are compact for all sufficiently large λ. We will show that for a { ∈ 0 ≤ } suitable choice of Ψ˜ 0 these sets are norm-bounded, and from this the Banach–Alaoglu theorem will give us compactness.

We take Ψ˜ 0 = (βj ,hj, ψj ) to have βj = 0 for all j. We take hj j such that Tj hj c0 > 0 a.e. on X, as guaranteed by the hypothesis of Theorem 4.1. By multiplying∈ Y by a suitable≥ positive d −1 constant if necessary, we can certainly assume that h < (2 G ′ ) . Finally, we take j=1 k jkYj k kX ψj = c01 which satisfies 0 ψj Tjhj . ≤ ≤ P For such a choice of Ψ˜ 0 we have d d d Λ((K,S ), Ψ˜ )= K 1 G ′ h + S ψ K/2+ c S 1. j 0  −k kX k jkYj  j j ≥ 0 j j=1 j=1 j=1 X X X Therefore, for given λ> 0,   (K,S ) C : Λ((K,S ), Ψ˜ ) λ [0, 2λ] S L∞(X)∗ : S(1) c−1λ d. { j ∈ j 0 ≤ }⊆ ×{ ∈ + ≤ 0 }

But it is easy to see that for S 0, S(1)= S ∞ ∗ . Indeed, by definition we have ≥ k kL (X) ∞ S ∞ ∗ = sup S(u) : u L (X), u 1 . k kL (X) {| | ∈ k k∞ ≤ } So let us take such a function u with u ∞ 1. Since S( u) = S(u) we may by choosing either u or u assume that S(u) 0. Wek k have≤ that u 1 a.e.− and therefore− the non-negativity of S gives us− that S(u) S(1), as≥ needed. ≤ ≤ Therefore (K,S ) C : Λ((K,S ), Ψ˜ ) λ [0, 2λ] S : S c−1λ d { j ∈ j 0 ≤ }⊆ ×{ k k≤ 0 } is a norm-bounded, weak-star closed, hence weak-star compact subset of R (L∞(X)∗)d, by the Banach–Alaoglu theorem. This completes the verification of condition (iv) of× Theorem 4.2 in our case, and we conclude that min sup L(Φ, Ψ) = sup inf L(Φ, Ψ). Φ∈C Ψ∈D Ψ∈D Φ∈C 5.2.2. Return to the main argument. We may therefore conclude, by Theorem 4.2, that if for non-zero G ′ fixed we define18 ∈ X ∗ γL = inf sup L and η = sup inf L (K,Sj )∈C (βj ,hj )∈D (βj ,hj )∈D (K,Sj )∈C

18 ∗ ∗ The notation here is perhaps confusing. We shall consider four problems: γ, γ , γL and γL. When there is no superscript we are dealing with the variant of the problem pertaining to L1, and presence of the superscript ∗ denotes that we are dealing with the variant of the problem which pertains to (L∞)∗; when there is no subscript we are dealing with the original version of the problem, and presence of the subscript L denotes that we are dealing with the Lagrangian formulation. This is consistent with the notation we adopted in the treatment of the finite discrete case above; in that case there was no distinction between L1 and (L∞)∗. We do not adorn η with either a superscript ∗ nor a subscript L since there is only one η-problem. Nevertheless we emphasise that the η-problem does indeed deal with the Langrangian formulation in the form pertaining to (L∞)∗. 30 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON

∗ ∗ then η = γL and the infimum in the problem for γL is achieved as a minimum. (It should be noted that we are not yet in a position to assert the finiteness of either of these numbers.)

In order to progress further, we shall also consider the problem γ = inf K d such that G(x) S (x)αj a.e. on X, and (40) ≤ j j=1 Y

S (x)T h (x) dµ(x) K G ′ h for all j and all h j j j ≤ k kX k jkYj j ∈ Yj ZX 1 where the Sj are taken to be in L (X, dµ). We emphasise that this is the problem we really want to solve: if we can prove that γ A and that minimisers exist, we will have our desired factorisation. Nevertheless, we should≤ point out that it is not yet even clear that there exist (K,Sj ) satisfying the constraints of (40). We shall be able to infer the existence of such (K,Sj ), and hence the finiteness of γ, only from the conclusion of Theorem 4.1. ∗ ∗ Our strategy is to show that (i) γL = γ and that if the problem for γL admits minimisers Φ, then the problem for γ also admits minimisers; and (ii) 0 η A. Combining these with the ∗ ∗ ≤ ≤ minimax result γL = η and existence of minimsiers for γL, we can conclude that the problem for γ admits minimisers and that γ A, which will conclude the proof of Theorem 4.1. ≤ ∗ ∗ Proof that γ = γL and that existence of minimisers for γL implies existence of min- ∗ imisers for γ. We begin by studying γL so let us consider for which (K,Sj ) C we have sup L< . Fix (K,S ). First of all, suppose that S are such that there exists∈ a tuple (β ) βj ,hj ∞ j j j such that d d G(x) βαj (x) dµ(x) α S (β ) > 0. j − j j j X j=1 j=1 Z Y X Then by setting hj = 0 and substituting βj tβj and letting t we see that the supremum is infinite. Therefore, if sup L< , we7→ must have → ∞ βj ,hj ∞ d d (41) G(x) βαj (x) dµ(x) α S (β ) for all simple functions β , j ≤ j j j j X j=1 j=1 Z Y X which, by Lemma 5.1, is equivalent to d (42) G(x) S (x)αj a.e. on X. ≤ jrn j=1 Y

Now assume there exists a j0 and an hj j such that Sj (Tj hj ) >K G X ′ hj Y . Taking 0 ∈ Y 0 0 0 0 k k k 0 k j0 βj = 0, and hj = 0 for j = j0 and multiplying hj0 by a factor t which we send to infinity we again see that the supremum6 is infinite. Therefore, if sup L< , then we must also have βj ,hj ∞ (43) S (T h ) K G ′ h j j j ≤ k kX k jkYj for all nonnegative hj and all j. From the positivity of Sjpfa we see that this implies

(44) S (x)T h (x) dµ(x) K G ′ h jrn j j ≤ k kX k j kYj ZX for all nonnegative hj and all j.

On the other hand, if for fixed (K,Sj) conditions (42) and (43) are satisfied, then when we are looking for supβj ,hj L, we can do no better than taking βj = 0 and hj = 0 for all j. So for MULTILINEAR DUALITY AND FACTORISATION 31

fixed (K,Sj ), we have supβj ,hj L< if and only if conditions (42) and (43) hold, in which case ∞∗ supβj ,hj L = K. So the problem for γL is identical with the problem γ∗ = inf K such that G(x) S (x)αj a.e., ≤ jrn j Y S (T h ) K G ′ h for all j and all h , j j j ≤ k kX k j kYj j ∈ Yj where we emphasise that the inf is taken over (K,S ) with S L∞(X)∗ . j j ∈ + Likewise, the problem

γL := inf sup L R 1 d (K,Sj )∈ +×(L (X)+) βj ,hj is identical with problem (40) for γ.

∗ It is clear that γL γL as the infimum for the left hand side is over a larger set than for the right hand side. ≤ Claim: γ γ∗ , and if minimisers Φ = (K,S ) exist for problem γ∗ , they also exist for problem L ≤ L j L γL. ∗ ∞ ∗ Indeed, assume that γL < , let ε > 0 and let (K,Sj ) with Sj L (X) and satisfying ∞ ∗ ∈ conditions (42) and (43) be such that K<γL + ε. Then the absolutely continuous component Sjrn satisfies (42) and (44), and so (K,Sjrn) contributes to the infimum in the problem for γL. ∗ Thus γL γL + ε. Letting ε 0 establishes the first part of the claim. Now suppose that ≤ → ∗ ∗ minimisers Φ = (K,Sj ) exist for problem γL. In particular this supposes that γL < . Let (K,S ) with S L∞(X)∗ and satisfying conditions (42) and (43) be such that K = γ∗∞. Then j j ∈ L the absolutely continuous component Sjrn satisfies (42) and (44), and so (K,Sjrn) contributes to and indeed achieves the infimum in the problem for γL (otherwise γL would be strictly less than ∗ γL). ∗ ∗ Summarising, the problems for γ and γL are equivalent; the problems for γ and γL are equivalent; ∗ ∗ γL = γL, and if extremisers exist for γL, they also exist for γL, and hence too for γ.

Proof that 0 η A. We wish to carry out a similar analysis for infK,Sj L, and for that we first of all rewrite≤ L≤as

d d αj L = G(x) β (x) dµ(x)+ K 1 G ′ h j  −k kX k jkYj  X j=1 j=1 Z Y X d   + S (T h α β ) . j j j − j j j=1 X We consider for which (β ,h ) D we have inf L> . j j ∈ K,Sj −∞

First, by taking Sj = 0 for all j =1,...,d and letting K to go infinity we see that if infK,Sj L> then we must have −∞ d (45) G ′ h 1. k kX k j kYj ≤ j=1 X Secondly, assume that there exists an index j0 and a set E X with µ(E) > 0 such that T h (x) < α β (x) for a.e. x E. Then by taking S = tχ⊆ L1, S = 0 for j = j and j0 j0 j0 j0 ∈ j0 E ∈ j 6 0 32 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON

K = 0 and letting t , then we see that infK,Sj L = . Thus if infK,Sj L> , we must also have → ∞ −∞ −∞ (46) α β (x) T h (x) a.e. on X for all j. j j ≤ j j If conditions (45) and (46) are both satisfied we can do no better than take K = 0 and Sj = 0 for all j. So, for fixed (βj ,hj ), infK,Sj L> if and only if conditions (45) and (46) hold, in which d αj −∞ case infK,Sj L = X G(x) j=1 βj (x)dµ(x). We can always find (βj ,hj ) such that conditions (45) and (46) hold, so η = sup G(x) d βαj (x)dµ(x) subject to conditions (45) and R Q βj ,hj X j=1 j (46). In particular this tells us that η 0. R≥ Q Let us now derive an upper bound for η. Examining the condition (46) on βj we see that

d α η sup G(x) α−1T h (x) j dµ(x) ≤ j j j hj ZX j=1 Y  d such that G ′ h 1. k kX k jkYj ≤ j=1 X Clearly there exist functions h such that G ′ h 1, and for any such we have j ∈ Yj k kX j k jkYj ≤ d P d −1 αj −1 αj G(x) α T h (x) dµ(x) G ′ T (α h )(x) j j j ≤k kX k j j j kX ZX j=1 j=1 Y  Y  d −1 αj G X ′ A α hj ≤k k k j kYj j=1 Y d −1 G ′ A α α h ≤k kX j k j j kYj j=1 X d = G ′ A h A k kX k jkYj ≤ j=1 X by H¨older’s inequality in the form Gf G X ′ f X , the multilinear inequality (15) which is our main hypothesis, the arithmetic-geometric≤k k meank k inequality and finally the assumption on R the hj . This clearly implies η A, and thus concludes the proof of Theorem 4.1. ≤ 

5.3. Consequences of saturation. We give two lemmas needed for Theorem 2.2. These allow us to construct suitable exhausting sequences of subsets of X of finite measure, in order that we might apply Theorem 4.1. Then we construct the weight w of the statement of Theorem 2.2.

Lemma 5.3. Let (X, dµ) be a σ-finite measure space, and suppose that (X)+ has the property that for every measurable set E X with µ(E) > 0, there existsP ⊆M an f and a subset E′ E with µ(E′) > 0, such that f⊆ > 0 a.e. on E′. Then there exists a countable∈ P subset ⊆ f N such that, with E := x X : f (x) > 0 , { n}n∈ ⊆P n { ∈ n } ∞ µ(X E )=0. \ n n=1 [ Proof. By exhausting X by a countable sequence of subsets, each of finite measure, we may assume that µ(X) is finite. We claim that for every ǫ> 0 there is a finite subset f ,...,f { 1 N }⊆P MULTILINEAR DUALITY AND FACTORISATION 33 such that N µ(X E ) < ǫ. \ n n=1 [ Once we have this claim, we take the union of the finite subsets of obtained for each ǫ =1/m, m N, and we are finished. P ∈ Suppose, for a contradiction, that there is some ǫ> 0 such that for all N, for all finite subfamilies f1,...,fN we have { }⊆P N µ(X E ) ǫ> 0. \ n ≥ n=1 [ Let N t = inf inf µ(X En). N {f1,...,fN }⊆P \ n=1 [ Then t ǫ> 0 and also t< since µ(X) is finite. For m N let m = f1,...,fN(m) be such that ≥ ∞ ∈ P { } N(m) µ(X E ) t +1/m; \ n ≤ n=1 [ we may assume that for all m. Letting m we obtain Pm ⊆Pm+1 → ∞ ∞ µ(X E ) t. \ n ≤ n=1 ∞ [ ∞ If µ(X n=1 En) = 0 we are done; otherwise E = X n=1 En has positive measure, and \ ′ \ ′ therefore, by hypothesis, there is a subset E E with µ(E )= δ > 0 such that for some f0 we have ES′ E . Then, (with the union now⊆ starting at n =S 0), ∈P ⊆ 0 N(m) µ(X E ) t +1/m δ. \ n ≤ − n=0 [ If we choose m>δ−1, we then have N(m) µ(X E ) < t, \ n n=0 [ in contradiction to the definition of t.  Lemma 5.4. Let (X, dµ) be a σ-finite measure space, a normed lattice, and suppose that T : (X) is a positive linear operator which saturatesY X. Then there is an increasing Y →M exhausting sequence of subsets (Gn) of X, each of finite measure, such that T strongly saturates each Gn. More precisely, there exists a sequence (hn) + such that hn+1 hn for all n, such that h 1 for all n, and such that for all n, Th ⊆(x Y) 1/n for x G .≥ k nkY ≤ n ≥ ∈ n Proof. Let = T ( +) (X)+. The saturation hypothesis allows us to deduce from the previous lemmaP thatY there⊆ M exists a sequence h such that if E = Th > 0 , then n ∈ Y+ n { n } E N covers X up to a set of measure zero. Letting { n}n∈ h h h h˜ =2−1 1 +2−2 2 + +2−n n , n h h ··· h k 1kY k 2kY k nkY we see that we may additionally assume that hn Y 1 for all n, and that hn+1 hn. Thus without loss of generality E E for all n k N.k ≤ ≥ n ⊆ n+1 ∈ 34 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON

Since X is σ-finite there is an increasing sequence of subsets Fn of finite measure which exhausts X. Now we set G := x : Th > 1/n F . n { n } ∩ n Clearly Gn Gn+1 for all n, and each Gn has finite measure. We check that Gn is exhausting. Let x X. Then⊆ x E for some k, i.e. Th (x) > 0, and therefore Th (x) >{1/l}for some l N. ∈ ∈ k k k ∈ Since X is exhausted by Fm there is some m such that x Fm. Therefore, for n such that n max k, l.m , we have{ that}x G . Finally, it is clear by definition∈ that T strongly saturates ≥ { } ∈ n Gn. 

As an immediate consequence, we have: Corollary 5.5. Let (X, dµ) be a σ-finite measure space and let be normed lattices for 1 Yj ≤ j d. Assume that Tj : j (X) for 1 j d are positive linear operators, each of which≤ saturates X. Then forY each→ M1 j d ≤there≤ exists a sequence (h ) such that ≤ ≤ j,n n ⊂ Yj hj,n Yj 1, hj,n hj,m for m n, and there exists an increasing and exhausting sequence of ksubsetsk E≤ X, each≤ of finite measure,≥ such that for each j and n, T h (x) 1/n for x E . n ⊆ j j,n ≥ ∈ n With this in hand, we can now define the weight w referred to in Remark 10 above. Let wj (x) for x E E be T h (x), where we take E = . Define w(x) = min w (x). Note that w is ∈ m \ m−1 j j,m 0 ∅ j j a.e. positive and a.e. finite. (If the sets Em stabilise in the sense that for some M N, EM = X up to a set of measure zero, then w 1/M, and we can simply take w to be 1). ∈ ≥ 5.4. Proof of Theorem 2.2. We will prove Theorem 2.2 by reducing it to Theorem 4.1. We will need the following lemma whose proof is an easy exercise in elementary point-set topology, and which is therefore omitted.

Lemma 5.6. Let Z be a compact topological space and suppose (zn) is an infinite sequence of distinct points in Z. Then there exists a point z Z such that every open neighbourhood of z ∈ contains infinitely many zn’s. Proof of Theorem 2.2. We may assume that A< otherwise there is nothing to prove. Take a ′ ∞ nonzero G , and take En as in Corollary 5.5. For each m we can apply Theorem 4.1, with X replaced∈ by XE , to conclude that there exist g L1(E , dµ) such that m j,m ∈ m d (47) G(x) g (x)αj a.e. on E , ≤ j,m m j=1 Y and such that for each j,

(48) g (x)T f (x)dµ(x) A G ′ f j,m j j ≤ k kX k j kYj ZEm for all f . j ∈ Yj If EM = X (up to a set of zero measure) for some M, we simply take gj = gj,M and we are finished. So we may assume that the sets Em do not stabilise, and therefore that there are infinitely many distinct gj,m for each j. With w defined as in the previous subsection, let us now calculate m

g 1 g (x) w (x) dµ(x)= g (x) T h (x) dµ(x) k j,mkL (w dµ) ≤ j,m j j,m j j,n Em n=0 En\En−1 Z X Z g (x) T h (x) dµ(x) A G ′ h A G ′ . ≤ j,m j j,m ≤ k kX k j,mkYj ≤ k kX ZEm MULTILINEAR DUALITY AND FACTORISATION 35

∞ ∗ Thus the functions gj,m all lie in a ball in L (X, w dµ) which, by the Banach–Alaoglu the- orem, is weak-star compact. It is therefore tempting to extract a weak-star convergent subse- quence. However, we must resist this temptation since L∞(X, w dµ) is not separable, and thus L∞(X, w dµ)∗ is not metrisable. We therefore proceed with some caution. We will use Lemma 5.6 as a substitute for the existence of weak-star convergent subsequences. It is convenient to consider the vectors g = (g ,...,g ) L1(X, w dµ) L1(X, w dµ) n 1,n d,n ∈ ×···× L∞(X, w dµ)∗ L∞(X, w dµ)∗ = (L∞(X, w dµ) L∞(X, w dµ))∗. ⊆ ×···× ×···× By Lemma 5.6 there is a point S = (S ,...,S ) (L∞(X, w dµ) L∞(X, w dµ))∗ such 1 d ∈ ×···× that every weak-star open neighbourhood of S contains infinitely many of the gn.

Lemma 5.7. Suppose (gn) and S are as above. (a) If for some q (X, w dµ)d we have ∈M + d g (q)= g q w dµ = g q w dµ K n j,n j n · ≤ j=1 X X X Z Z for all sufficiently large n, then S(q) K. ≤ (b) If for some q L∞(X, w dµ)d we have ∈ d g (q)= g q w dµ = g q w dµ L n j,n j n · ≥ j=1 X X X Z Z for all sufficiently large n, then S(q) L. ≥ Proof. (a) Suppose for a contradiction that S(q) K′ >K for some finite K′. Let ≥ U = R ((L∞(w dµ))d)∗ : R(q) > (K + K′)/2 . { ∈ } d Then S U, and U is weak-star open since for each q (w dµ)+ the functional R R(q) is weak-star∈ lower semicontinuous, by (a vector-valued version∈M of) Lemma 5.2. Thus U 7→is an open neighbourhood of S in the weak-star topology. By the above remarks, U must contain infinitely many of the (gn). But for all n sufficiently large,

g (q)= g q w dµ K < (K + K′)/2 n n · ≤ ZX and so none of these g can be in U. This is a contradiction, and therefore S(q) K. n ≤ (b) Suppose for a contradiction that S(q)= L′

g (q)= g q w dµ L> (L + L′)/2 n n · ≥ ZX and so none of these g can be in U. This is a contradiction, and therefore S(q) L.  n ≥ 36 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON

We now wish to verify that the absolutely continuous components (Sjrn) (where the Radon– Nikodym derivative is with respect to the measure w dµ) of (Sj ) satisfy

d G(x) S (x)αj a.e. on X, and ≤ jrn j=1 (49) Y

S (x)T f (x) dµ(x) A G ′ f jrn j j ≤ k kX k jkYj ZX ∞ ∗ for all j and for all fj j . Since we know that Sj (L (X, w dµ)) , we will therefore have S L1(w dµ), and this∈ Y will conclude the proof of Theorem∈ 2.2. jrn ∈ We may suppose that G ′ = 1. k kX We look at the second inequality from (49) first. Fix m and consider

−1 Sjrn(x)Tj fj (x)dµ(x)= Sjrn(x)w(x) χEm (x)Tj fj(x)w(x)dµ(x) ZEm ZX S (w−1χ T f ) ≤ j Em j j by positivity of each component in the Yosida–Hewitt decomposition of Sj , (recall Theorem 4.4). Now, for n m, ≥ g (x)[w(x)−1T f (x)χ (x)]w(x)dµ(x) g (x)T f (x)dµ(x) A f , jn j j Em ≤ jn j j ≤ k j kYj Z Z so that by Lemma 5.7(a) (in the scalar case), S (w−1T f χ ) A f . j j j Em ≤ k jkYj Thus S (x)T f (x)dµ(x) A f jrn j j ≤ k jkYj ZEm and we now let m to get the second inequality of (49). → ∞ Now we look at the first inequality from (49). By Lemma 5.1 (using the measure w dµ) and the fact that the Em exhaust X, it suffices to show that for each fixed m, and all simple βj ,

d d G(x) β (x)αj w(x)dµ(x) α S (β ). j ≤ j j j Em j=1 j=1 Z Y X Take n m.By (47) and the arithmetic-geometric mean inequality, the left-hand side is at most ≥ d d [g (x)β (x)]αj w(x)dµ(x) α g (x)β (x)w(x)dµ(x) jn j ≤ j jn j Em j=1 j=1 Em Z Y X Z d d α g (x)β (x)w(x)dµ(x) = g (x)[α β (x)]w(x)dµ(x). ≤ j jn j jn j j j=1 X j=1 X X Z X Z Thus for all n m, ≥ d d g (x)[α β (x)]w(x)dµ(x) G(x) β (x)αj w(x)dµ(x). jn j j ≥ j j=1 X Em j=1 X Z Z Y MULTILINEAR DUALITY AND FACTORISATION 37

Since the simple functions βj are bounded, Lemma 5.7(b) gives us that d d S (α β ) G(x) β (x)αj w(x)dµ(x), j j j ≥ j j=1 Em j=1 X Z Y which is what we want. This completes the proof of Theorem 2.2. 

Part II. Connections with other topics

6. Complex interpolation and factorisation We begin by observing that the trivial identity of Example 2,

f1(x2)f2(x1) dx1dx2 = f1 f2, R2 R R Z Z Z immediately implies via Theorem 1.3 that, for every nonnegative G L2(R2), there exist non- ∈ negative g1 and g2 such that G(x) g (x)g (x) for almost every x R2 ≤ 1 2 ∈ and p ess sup g (x , x )dx G and ess sup g (x , x )dx G . x2 1 1 2 1 ≤k k2 x1 2 1 2 2 ≤k k2 Z Z While it is not perhaps entirely obvious how to do this explicity (a point to which we return in Sections 9.2, 9.3 and 10.2.1 below), for now we want to point out that this example highlights the connection between our multilinear duality theory and the theory of interpolation of Banach spaces. In particular, we consider the upper method of complex interpolation of A. P. Calder´on, [19].

Suppose that Z0 and Z1 are Banach lattices of measurable functions defined on some measure space. We define Z1−θZθ = f : there exist f Z such that f f 1−θ f θ 0 1 { j ∈ j | | ≤ | 0| | 1| } with 1−θ θ 1−θ θ f Z Z = inf f0 Z f1 Z1 , k k 0 1 {k k 0 k k } the inf being taken over all possible decompositions of f. Under the assumption that the unit 1−θ θ ball of Z0 Z1 is closed in Z0 + Z1, Calder´on showed that 1−θ θ θ Z0 Z1 = [Z0,Z1] θ where [Z0,Z1] is the interpolation space between Z0 and Z1 obtained by the upper complex method. With this in mind, the factorisation statement in our example is tantamount to the statement .L2(R2) ֒ [L∞ (L1 ),L∞ (L1 )]1/2 → x1 x2 x2 x1 Many other special cases of our theory can be similarly expressed in the language of interpolation. We leave it to the interested reader to pursue this point of view more systematically. In this particular example, there is further structure, see for example Pisier [37], (in which some of the ideas are attributed to Lust-Piquard). There it is established that we have L2(R2)= HS(L2(R)) ֒ (L2) = [L∞ (L1 ),L∞ (L1 )]1/2 → Lreg x1 x2 x2 x1 38 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON

2 where HS denotes the class of Hilbert–Schmidt operators and reg(L ) is the space of regular bounded linear operators on L2. In rough terms, a regular boundedL linear operator on L2 is one such that if its kernel is K(s,t), then K(s,t) is also the kernel of a bounded linear operator. | | The implicit factorisation arguments involved in establishing results of this type rely on the Hahn– Banach theorem or the Perron–Frobenius theorem, and are thus related to minimax theory; they are similarly non-constructive.

7. Factorisation and convexity It is also natural to enquire about how factorisation and interpolation interact at the level of particular families of inequalities. For the sake of concreteness, suppose we are in the setting of multilinear generalised Radon transforms on euclidean spaces – so that T F = F B for j j j ◦ j suitable Bj . We shall suppress consideration of any of the technical hypotheses of Theorem 2.2 in what follows. Suppose that we have the pair of inequalities

d d

(50) T F A F p j j ≤ k k j kL jk j=1 q j=1 Y L k Y for k =0, 1, where qk,pjk 1. ≥ Each of these has a family of corresponding equivalent factorisation statements, according to Theorem 2.2 and the remarks in Section 1.3. See also Section 7.1 below. After some changes of −1 notation, one such equivalent pair of statements is as follows. For k =0, 1, let sk := qk j pjk . ′ sk Then for all nonnegative Gk (k = 0, 1) such that Gk = 1, there are nonnegative g10,...,gP d0 and g ,...,g such that 11 d1 R d (51) G (x) g (x)qk /pjk sk a.e. k ≤ jk j=1 Y and such that for all f with f 1, j j ≤ (52) R f (B x)g (x)dx Aqk /sk j j jk ≤ k Z for k =0, 1. From (51) and (52) we shall deduce a factorisation statement which implies the natural interpo- lation statement d d 1−θ θ (53) T F A A F p j j ≤ 0 1 k j kL jθ j=1 q j=1 Y L θ Y for 0 <θ< 1, where, as usual, 1/qθ = (1 θ)/q0 + θ/q1, and similarly for 1/pjθ. − 1/s′ Indeed, given a nonnegative G with G = 1, let Gk = G k . Taking convex combinations in (51) gives us R d ′ ′ (54) G(x) g (x)q0s0(1−θ)/pj0s0 g (x)q1s1θ/pj1s1 a.e. ≤ j0 j1 j=1 Y Next, we define ′ ′ q0s0 q1s1 γj (θ) := (1 θ)+ θ pj0s0 − pj1s1 MULTILINEAR DUALITY AND FACTORISATION 39 and define gjθ by γ (θ) ′ ′ j q0 s0(1−θ)/pj0s0 q1 s1θ/pj1s1 gjθ := gj0(x) gj1(x) . Then, by (52), we have

′ ′ q0 s (1−θ)/pj0s0γj (θ) q1 s θ/pj1s1γj (θ) fj (Bj x)gjθ(x)dx = fj (Bj x)gj0(x) 0 gj1(x) 1 dx Z Z ′ ′ q0s0(1−θ)/pj0s0γj (θ) q1s1θ/pj1s1γj (θ) f (B x)g (x)dx f (B x)g (x)dx ≤ j j j0 j j j1 Z  Z  by H¨older’s inequality, since γj(θ) is defined precisely to ensure the two exponents on the right hand side here add to 1. Therefore, if f 1, j ≤ ′ ′ R q0s0(1−θ)/pj0s0γj (θ) q1s1θ/pj1s1γj (θ) f (B x)g (x)dx Aq0/s0 Aq1/s1 . j j jθ ≤ 0 1 Z h i h i d Now let βj (θ) := λ(θ)γj (θ) where λ(θ) is defined so that j=1 βj (θ) = 1. By the definition of s0 and s1 we have P d d ′ ′ q0s0 q1s1 ′ ′ γj (θ)= (1 θ)+ θ = (1 θ)s0 + θs1. p 0s0 − p 1s1 − j=1 j=1 j j X X   So, we take 1 λ(θ) := . (1 θ)s′ + θs′ − 0 1 Now, bearing in mind Remark 7, we conclude that

d βj (θ) ′ ′ Pj q0s0(1−θ)βj (θ)/pj0s0γj (θ) Pj q1s1θβj (θ)/pj1s1γj (θ) f (B x)g (x)dx Aq0/s0 Aq1/s1 j j jθ ≤ 0 1 j=1 Y Z  h i h i ′ ′ ′ ′ λ(θ) Pj q0s0(1−θ)/pj0s0 λ(θ) Pj q1s1θ/pj1s1 λ(θ)s0(1−θ) λ(θ)s1θ q0/s0 q1/s1 q0/s0 q1 /s1 = A0 A1 = A0 A1 h i h i h i h i Q(θ)/S(θ) ′ ′ ′ ′ S(θ)/Q(θ) s0q0λ(θ)(1−θ) s1q1λ(θ)θ s0q0λ(θ)(1−θ) s1q1λ(θ)θ s0 s1 s0 s1 = A0 A1 = A0 A1 " #    for a certain quantity S(θ)/Q(θ) to which we turn our attention next. Indeed we define this quantity (not S(θ), Q(θ) separately), so that the exponents on A0 and A1 inside the curly brackets sum to 1. That is, Q(θ) s′ q (1 θ) s′ q θ := λ(θ) 0 0 − + 1 1 . S(θ) s s  0 1  Let us define these exponents of A0 and A1 as 1 α(θ) and α(θ) respectively; that is, we define α(θ) by − S(θ) s′ q θ α(θ) := λ(θ) 1 1 . Q(θ) s1 40 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON

Next, we want the β = λγ to be of the form β (θ) = Q(θ) for certain P (θ); that is, j j j Pj (θ)S(θ) j 1 = S(θ)βj (θ) = S(θ)λ(θ)γj (θ) . So, bearing in mind the definitions of γ and S/Q, we define Pj (θ) Q(θ) Q(θ) j Pj (θ) by ′ ′ q0s q1s 1 0 (1 θ)+ 1 θ pj0s0 − pj1s1 := s′ q (1−θ) s′ q θ . Pj (θ) 0 0 + 1 1 s0 s1 Finally, we define Q(θ) by 1 1 1 := (1 α(θ)) + α(θ) . Q(θ) − q0 q1 It is not hard to check that with all these definitions in place, we have, for each j, 1 1 1 = (1 α(θ)) + α(θ) . Pj (θ) − pj0 pj1

′ ′ We therefore have that for each 0 θ 1, for all G = G1/S (θ) such that GS (θ) = 1, there ≤ ≤ θ θ exist gjθ such that d R G (x) g (x)Q(θ)/Pj (θ)S(θ) θ ≤ jθ j=1 Y and, for f such that f 1, j j ≤ d R Q(θ)/Pj (θ)S(θ) Q(θ)/S(θ) f (B x)g (x)dx A1−α(θ)Aα(θ) . j j jθ ≤ 0 1 j=1 Y Z    d Note particularly that the exponents Q(θ)/Pj (θ)S(θ) sum to 1 since j=1 βj = 1. Consequently, using the flexibility that Remark 7 affords us, P d d 1−α(θ) α(θ) T F A A F P (θ) j j ≤ 0 1 k j kL j j=1 j=1 Y LQ(θ) Y for 0 <θ< 1. Noting that the map α : [0, 1] [0, 1] is a surjection completes the argument proving (53). → The argument given here provides no insight into cases in which (53) might hold with a smaller 1−θ θ constant than A0 A1. 7.1. Factorisation and multiple manifestations of generalised Radon transforms. As we have observed in Section 1.3 there may be multiple equivalent manifestations of the same multilinear inequality. For concreteness, suppose that we are once again considering multilinear generalised Radon transforms on euclidean spaces so that Tj f = f Bj for suitable Bj . Then the two inequalities ◦ d d αj αj (Tj fj ) A fj ≤ pj j=1 j=1 Y q Y and d d α˜j α˜j (Tj f˜j ) A˜ f˜j ≤ p˜j j=1 j=1 Y q˜ Y d d (where we are imposing α =1= α˜ ) are clearly equivalent provided that Aq = A˜q˜ j=1 j j=1 j and α p˜ /α˜ p =q/q ˜ for all j. The corresponding factorisation statements j j j j P P MULTILINEAR DUALITY AND FACTORISATION 41

′ For all nonnegative G Lq there exist nonnegative locally integrable functions g such that ∈ j d G(x) g (x)αj a.e. ≤ j j=1 Y and such that for each j, for all f Lpj , j ∈

g (x)f (B x)dx A G ′ f . j j j ≤ k kq k j kpj Z and

′ For all nonnegative G˜ Lq˜ there exist nonnegative locally integrable functions g˜ such that ∈ j d G˜(x) g˜ (x)α˜j a.e. ≤ j j=1 Y and such that for each j, for all f˜ Lp˜j , j ∈

g˜ (x)f˜ (B x)dx A˜ G˜ q˜′ f˜ . j j j ≤ k kL k j kp˜j Z are therefore also equivalent (subject to suitable hypotheses), by Proposition 1.1 and Theo- rem 1.3. However it is not immediately apparent whether this equivalence can be seen directly via changes of notation coupled with simple convexity arguments. In this connection the remarks in Section 5.7 of [17] may be helpful.

8. Factorisation and more general multilinear operators The multilinear operators we have considered have a rather special form in so far as they are built out of a collection of positive linear operators by taking a pointwise geometric mean. One may ask to what extent the theory we have developed is valid for more general multilinear operators T : 1 d . In such a setting we will no longer be able to attribute different “weights” α toY the×···×Y different→ components X , and all of them will need to be treated on an equal footing. j Yj For a nonnegative kernel K, let us therefore consider multilinear operators of the form

T (f ,...,f )(x)= K(x, y ,...,y )f (y ) ...f (y )dν (y ) ... dν (y ) 1 d ··· 1 d 1 1 d d 1 1 d d ZYd ZY1 and inequalities of the form

d 1/d 1/d (55) T (f1,...,fd) A fj Y . X ≤ k k j j=1 Y

When K is of the form K(x, y1,...,yd) = K1(x, y1) ...Kd(x, yd), these are the special cases of the inequalities (15) given by αj = 1/d for j = 1,...,d. It is very natural to ask whether there is a general duality/factorisation result along the same lines as Proposition 2.1 and Theorem 2.2 which yields a necessary and sufficient condition for the validity of inequality (55). As the reader will readily verify by following the proof of Proposition 2.1, inequality (55) does indeed hold (under hypotheses on and similar to those of Proposition 2.1), if, for all G ′ X Yj ∈ X 42 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON such that G X ′ 1, we have that there there exist nonnegative functions gj on X Yj such that k k ≤ × d K(x, y ,...,y )1/dG(x) g (x, y )1/d a.e. 1 d ≤ j j j=1 (56) Y

and gj (x, )dµ(x) A. X · Y∗ ≤ Z j

This observation has proved very useful in multilinear Kakeya theory, see Section 11 below. However, the converse is not true, namely inequality (55) does not in general imply the existence of Sj such that (56) holds even if we assume that the integral kernel K is invariant under permutations of the y-variables:

4 Proposition 8.1. Let d =2. Let X = Y1 = Y2 = 1, 2 =Ω with counting measure, = L (Ω), and = = L2(Ω). There exists a bilinear T :{L2(Ω)} L2(Ω) L4(Ω) such thatX(55) holds Y1 Y2 × → with A =21/4 but such that (56) can only hold with A 21/2. ≥ Proof. Let the integral kernel K of T satisfy K(1, 1, 1) = K(2, 1, 1) = K(2, 2, 2)=1 2 2 and let K equal zero otherwise. Let f1 = (a1,a2) and f2 = (b1,b2) and we assume a1 + a2 = 2 2 b1 + b2 = 1. Then

T (f1,f2)(1) = a1b1 and T (f1,f2)(2) = a1b1 + a2b2 so 2 2 2 T (f1,f2)(x) dx = (a1b1) + (a1b1 + a2b2) . Z This is clearly maximised, subject to the normalisation constraints, by taking a1 = b1 = 1, a2 = b2 = 0 and the maximum is 2. So we see that the multilinear inequality (55) holds for this operator with A =21/4. For problem (56), consider G = (0, 1). Then the non-trivial constraints are 1 g (2, 1)g (2, 1) and 1 g (2, 2)g (2, 2) ≤ 1 2 ≤ 1 2 and p p (g (1, 1)+ g (2, 1))2 + (g (1, 2)+ g (2, 2))2 A 1 1 1 1 ≤ p(g (1, 1)+ g (2, 1))2 + (g (1, 2)+ g (2, 2))2 A. 2 2 2 2 ≤ Using uv (u2 + v2)/2p on the lower bounds gives ≤ 1 (g (2, 1)2 + g (2, 1)2)/2 and 1 (g (2, 2)2 + g (2, 2)2)/2, ≤ 1 2 ≤ 1 2 so 2 g (2, 1)2 + g (2, 1)2 and 2 g (2, 2)2 + g (2, 2)2, ≤ 1 2 ≤ 1 2 so 4 g (2, 1)2 + g (2, 1)2 + g (2, 2)2 + g (2, 2)2, ≤ 1 2 1 2 and thus 2 max g (2, 1)2 + g (2, 2)2,g (2, 1)2 + g (2, 2)2 ≤ { 1 1 2 2 } giving A 21/2, which is strictly larger than 21/4. So while inequality (55) holds in this case, ≥ there are G for which there are no gj satisfying (56) with the same value of A.  MULTILINEAR DUALITY AND FACTORISATION 43

We invite the reader to use this idea to construct examples where (55) holds with A = 1 but for which (56) holds for no finite A. See [28] for a different approach to inequalities of the form (55), based upon considerations related to Schur’s lemma rather than duality.

Part III. Examples and illustrations of the theory

In this part we revisit the examples in the introduction which motivated our study. We examine what insights our duality–factorisation results bring to, and have gained from, each of them. In some cases we reap the benefits of more direct and streamlined factorisation-based proofs of known inequalities. In others, an interesting challenge is posed – it can be argued that we cannot claim to have a full understanding of an inequality until we can exhibit its equivalent factorisation statement.

9. Classical inequalities revisited 9.1. H¨older’s inequality. We observed above that the multilinear form of H¨older’s inequality d for nonnegative functions is equivalent, for any fixed set of exponents αj > 0 with j=1 αj = 1, to α1 αd α1 αd P f f q f1 fd k 1 ··· d k ≤k kq1 ···k kqd for any choice of indices 1 q < and 1 q< which satisfies d α q−1 = q−1. ≤ j ∞ ≤ ∞ j=1 j j By Theorem 2.2, each instance of this inequality is equivalent to the existenceP of a subfactorisation ′ of any G Lq as ∈ d G(x) g (x)αj a.e. ≤ j j=1 Y where

g ′ G ′ . k jkqj ≤k kq Taking g = λ Gγj for appropriate λ and γ verifies this. In particular, if we take q = q 1 j j j j j ≥ for all j, then we can simply take gj = G for all j.

9.2. The affine-invariant Loomis–Whitney inequality. Recall that the Loomis–Whitney inequality is

F1(π1x) Fn(πnx) dx F1 Ln−1(Rn−1) Fn Ln−1(Rn−1), | Rn ··· |≤k k ···k k Z where πj x = (x1,..., xj ,...,xn) is projection onto the hyperplane perpendicular to the j’th standard basis vector e . For every 0

More generally, if πω⊥ represents orthogonal projection onto the hyperplane perpendicular to ω Sn−1, we have the affine-invariant Loomis–Whitney inequality ∈ n n 1/(n−1) 1/(n−1) −1/(n−1) (57) fj (πω⊥ x) dx (ω1 ωn) fj , Rn j ≤ ∧···∧ Rn−1 j=1 j=1 Z Y Y Z  where (ω ω )−1/(n−1) is the best constant in the inequality. Here, ω ω is the 1 ∧···∧ n 1 ∧···∧ n modulus of the determinant of the matrix whose columns are ω1,...,ωn, and it is the volume of the parallepiped whose sides are given by the vectors ωj. (Clearly if we choose all the ωj to be the same we cannot expect a finite constant, and the constant in general should reflect “quantitative linear independence” of the ωj .) We give a direct and elegant proof of (57) by explicitly establishing a suitable factorisation. Indeed, according to Proposition 1.1, it is sufficient that for every nonnegative G Ln(Rn) we ∈ can find g1,...,gn such that G(x)= g (x)1/n g (x)1/n a.e. 1 ··· n and, for all j and almost every x,

g (x + tω )dt = (ω ω )−1/n G . j j 1 ∧···∧ n k kn Z This is because for any f : Rn−1 R and g : Rn R, writing x Rn as x = u + tω with → → ∈ j u ω⊥, we have ∈ j

f(πω⊥ x)g(x)dx = f(πω⊥ (u + tωj ))g(u + tωj )dtdu j Rn−1 R j Z Z Z

= f(u) g(u + tωj )dt du. Rn−1 R Z Z  Let G : Rn R be a nonnegative function which satisfies G(x)ndx = 1. For ω ,...,ω → Rn 1 n ∈ Sn−1 and ξ Rn let us first note that if we set, for s = (s ,...,s ) Rn, ∈ 1 R n ∈ y(s)= ξ + s ω + + s ω + s ω , 1 1 ··· n−1 n−1 n n then we have that the Jacobian map ∂y/∂s satisfies det (∂y/∂s) = ω ω . | | 1 ∧···∧ n Therefore, for every ξ Rn, ∈ G(ξ + s ω + + s ω + s ω )nds ds ... ds = G(y(s))nds 1 1 ··· n−1 n−1 n n 1 2 n Z Z 1 = G(y)n dy = (ω ω )−1. det (∂y/∂s) 1 ∧···∧ n Z | | Secondly, G(x)n can be written (for a.e. x) as a telescoping product G(x)n G(x + s ω )nds 1 1 1 ... G(x + s ω )nds G(x + s ω + s ω )nds ds 1 1 1 × R 1 1 2 2 1 2 × G(x + s ω + + s ω )nds ds ... ds R 1 1 ···R n−1 n−1 1 2 n−1 (ω ω )−1 × G(x + s ω + + s ω + s ω )nds ds ... ds × 1 ∧···∧ n R 1 1 ··· n−1 n−1 n n 1 2 n R := g1(x) ...gn(x) MULTILINEAR DUALITY AND FACTORISATION 45 where G(x + s ω + + s ω )nds ds ... ds g (x)= 1 1 ··· j−1 j−1 1 2 j−1 (ω ω )−1/n . j G(x + s ω + + s ω + s ω )nds ds ... ds × 1 ∧···∧ n R 1 1 ··· j−1 j−1 j j 1 2 j If we replace Rx by x + tωj in this formula, the denominator is unchanged, and so if we then integrate with respect to t we immediately see that

g (x + tω )dt = (ω ω )−1/n j j 1 ∧···∧ n Z identically for x Rn, as we needed. ∈ A similar approach works when we instead consider projections onto subspaces whose codi- n mensions sum to n. Indeed, suppose that we have subspaces Ej of R with dim Ej = kj and d k = n and assume that Rn = E + + E as an algebraic direct sum. j=1 j 1 ··· d WeP identify a quantity which measures lack of orthogonality of these subspaces in the same way n−1 that the wedge product ω1 ωn measures the degeneracy in the directions ω1,...,ωn S . Let e ,e ,...,e be an∧···∧ orthonormal basis for E and define ∈ { j1 j2 jkj } j E E := d kj e ; 1 ∧···∧ d ∧j=1 ∧k=1 jk that is, E E is the absolute value of the determinant of the n n matrix whose j’th 1 ∧···∧ d × block of kj columns comprises an orthonormal basis for Ej . It is easily checked that this quantity is independent of the particular orthonormal bases chosen, and it can of course be defined in a more canonical and invariant way.

Proposition 9.1. For Ej as above, let πj be the projection whose kernel is Ej . Then we have the affine-invariant kj -plane Loomis–Whitney inequality:

1/(d−1) 1/(d−1) f1(π1x) ...fd(πdx) dx Rn Z (58) 1/(d−1) 1/(d−1) (E E )−1/(d−1) f ... f . ≤ 1 ∧···∧ d 1 d Z  Z 

The proof via factorisation is formally the same as the case when kj = 1 for all j, where now the roles of the variables s R1 are replaced by copies of Rkj . We leave the details to the reader. j ∈ In the special case of the trivial identity,

F1(x2)F2(x1) dx1dx2 = F1 F2, R2 R R Z Z Z (see Section 6), a suitable factorisation of G L2(R2) with G = 1 is given by G(x)2 = ∈ k k2 g1(x)g2(x) a.e., where 2 G(x1, x2) g1(x1, x2)= 2 R G(s, x2) ds and R 2 g2(x1, x2)= G(s, x2) ds. R Z Note that this factorisation depends upon the order we have assigned to 1, 2 . On the other hand, given this ordering, the essentially unique way to write { } 2 G(x) = g1(x1, x2)g2(x2) where g1( , x2) 1 = 1 for all x2 and g2 1 = 1 is as we have given. See Section 10.2.1, where this observationk · k drives related issues. k k 46 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON

There are many variants of the Loomis–Whitney inequality – for example Finner’s inequalities [25] – which can likewise be established by the same factorisation method. 9.3. The nonlinear Loomis–Whitney inequality. Nonlinear Loomis–Whitney inequalities (and some multilinear generalised Radon transforms) can likewise be established by similar meth- ods. In fact the first proof of the nonlinear Loomis–Whitney inequality with essentially the sharp constant was obtained via an explicit factorisation technique. We give the details. Let V be an open neighbourhood of 0 in Rn and U an open neighbourhood of 0 in Rn−1. Let π : V U be a C1 submersion onto U, and for x V let ω(x) be the wedge product of the rows of→ dπ(x). We assume that the fibres π−1(u) for∈u U can be parametrised by C1 curves t γ(t, x) in such a way that ∈ 7→ for all x V , γ(0, x)= x • for all x ∈ V , for all t, πγ(t, x)= πx • (semigroup∈ property) for all x V , for all t and s, • ∈ γ(t,γ(s, x)) = γ(s + t, x) for all x and t, d γ(t, x)= ω(γ(t, x)). • dt The domain of each curve γ( , x) will be an open interval I containing 0 which we largely · x suppress in what follows, but we stress that γ(Ix, x) is the entire fibre containing x. In all the t-integrals below it is assumed that we are integrating over such maximal domains. We note that under these assumptions, especially the last one, the co-area formula gives

f(πx)g(x)dx = f(u) g(γ(t, u˜))dt du ZV ZU Z  for any reasonable functions f and g.

We now assume that we have n submersions π1,...,πn as above, and we assume that ω1(0) ω (0) = 0. For each x V we define the maps t Φ (t) by ∧ ···∧ n 6 ∈ 7→ x Φ : (t ,...,t ) γ (t ,γ (t ,...,γ (t , x)) ... ) x 1 n 7→ 1 1 2 2 n n which satisfy Φx(0) = x and also det(DΦ )(0) = (ω ω )(x) =0 | x | 1 ∧···∧ n 6 provided x is sufficiently close to 0.

We shall assume that V is sufficiently small so that for each x V , the map Φx is injective – as was pointed out in [11], even in two dimensions some global hypothesis∈ of this sort is needed. With the set-up above, for x V let ∈ −1 (59) W (x) := inf det (DΦξ)(Φ (x)) . ξ∈V | ξ |

Note that W (x) ω1(x) ωn(x), (take ξ = x), and that W (Φx(t)) det(DΦx)(t) for all x and t. ≤ ∧···∧ ≤ | | For 1 j n and suitable F let ≤ ≤ F (γ (t ,γ (t ,...,γ (t , x)) ... ))dt ... dt S (x)= ··· 1 1 2 2 j−1 j−1 j−1 1 j F (γ (t ,γ (t ,...,γ (t , x)) ... ))dt ... dt R R··· 1 1 2 2 j j j 1 (so that S1 has no integralsR in theR numerator). Then we have F (γ (t ,γ (t ,...,γ (t ,γ (τ, x))) ... ))dt ... dt S (γ (τ, x)) = ··· 1 1 2 2 j−1 j−1 j j−1 1 . j j F (γ (t ,γ (t ,...,γ (t ,γ (τ, x))) ... ))dt ... dt R R··· 1 1 2 2 j j j j 1 R R MULTILINEAR DUALITY AND FACTORISATION 47

We claim that for each j and each x,

Sj (γj (τ, x))dτ =1. Z Indeed, notice that the denominator in the previous expression,

F (γ (t ,γ (t ,...,γ (t ,γ (τ, x))) ... ))dt ... dt , ··· 1 1 2 2 j j j j 1 Z Z equals F (γ (t ,γ (t ,...,γ (t + τ, x)) ... ))dt ... dt ··· 1 1 2 2 j j j 1 Z Z = F (γ (t ,γ (t ,...,γ (t , x)) ... ))dt ... dt ··· 1 1 2 2 j j j 1 Z Z by the semigroup property, and is therefore independent of τ. So F (γ (t ,γ (t ,...,γ (t ,γ (τ, x))) ... ))dt ... dt dτ S (γ (τ, x))dτ = ··· 1 1 2 2 j−1 j−1 j j−1 1 j j F (γ (t ,γ (t ,...,γ (t , x)) ... ))dt ... dt Z R R ··· 1 1 2 2 j j j 1 which equals 1 by Fubini’s theorem.R R On the other hand, n F (x) S (x)= , j F (γ (t ,γ (t ,...,γ (t , x)) ... ))dt ... dt j=1 1 1 2 2 n n n 1 Y ··· so that R R n F (x)= Sj (x) F (Φx(t))dt. j=1 Y Z Taking F (x)= S(x)nW (x), we therefore have n n n S(x) W (x)= Sj (x) S(Φx(t)) W (Φx(t))dt j=1 Y Z n n S (x) S(Φ (t))n det(DΦ )(t) dt = S (x) S(y)ndy ≤ j x | x | j j=1 j=1 V Y Z Y Z since W (Φx(t)) det(DΦx)(t) for all t and since each Φx is injective. We also have that for each j and each ≤x, | |

f(πj x)Sj (x)dx = f(u) Sj (γj (τ, u˜))dτ du = f. ZV ZUj Z  ZUj 1 By the easy half of the duality argument, this shows that for all nonnegative fj L (Uj ) we have ∈ n n 1/n 1/n 1/n f (π x) W (x) n/(n−1) f . k j j kL (V ) ≤ j j=1 j=1 Uj ! Y Y Z Consequently we have: Proposition 9.2. Under the above assumptions, with W defined as in (59), we have

n n 1/(n−1) f (π x)1/(n−1)W (x)1/(n−1) dx f . j j ≤ j V j=1 j=1 Uj ! Z Y Y Z 48 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON

Noting that W (x) ω (x) ω (x), one might ask whether ≤ 1 ∧···∧ n n n 1/(n−1) f (π x)1/(n−1)ω (x) ω (x)1/(n−1) dx f j j 1 ∧···∧ n ≤ j V j=1 j=1 Uj ! Z Y Y Z holds for sufficiently small V . As an immediate corollary of Proposition 9.2, we obtain the sharp form of the nonlinear Loomis– Whitney inequality of [11]: Corollary 9.3. Let V be an open neighbourhood of 0 in Rn and U an open neighbourhood of 0 in Rn−1. For 1 j n, let π : V U be C1 submersions onto U, and for x V let ω (x) be ≤ ≤ j → ∈ j the wedge product of the rows of dπj . Assume that ω1(0) ωn(0) =0. Then, for all ǫ> 0, there is a neighbourhood V ′ V of 0, such that for all f ∧···∧ 6 ⊆ j n n 1/(n−1) 1/(n−1) −1/(n−1) fj (πj x) dx (1 + ǫ)ω1(0) ωn(0) fj ′ ≤ ∧···∧ V j=1 j=1 Uj ! Z Y Y Z ′ n−1 Proof. Given ǫ> 0 we can choose V sufficiently small that ω1(0) ωn(0) (1 + ǫ) W (x) for all x V ′. ∧···∧ ≤  ∈ Since the work in this section was presented in various public fora, Bennett et al have shown, using methods based on induction on scales, that any Brascamp–Lieb inequality has a corresponding nonlinear counterpart with the same loss in the constant of at most (1 + ǫ). See [7].

10. Brascamp–Lieb inequalities revisited We shall discuss the Brascamp–Lieb inequalities under two headings. Firstly we shall address geometric Brascamp–Lieb inequalities (where in particular we can identify the sharp constant and existence of Gaussian extremisers), and secondly we will examine general Brascamp–Lieb inequalities with a finite (but unquantified) constant. 10.1. Geometric Brascamp–Lieb inequalities. The next result is a direct application of Theorem 1.3 to the geometric Brascamp–Lieb inequalities of Example 3. n n Theorem 10.1. For 1 j d let Vj be a subspace of R . Let Bj : R Vj be orthogonal projection. Suppose there≤ exist≤p with 0

gj G q′ . ⊥ ′ ≤k k Vj qj Z L (Vj )

One simply needs to note (see the discussion in Example 3) that under the hypothesis of this theorem, d d pj qj /q pj qj /q (fj Bj ) fj qj ◦ ≤ k kL (Vj ) j=1 j=1 Y Lq(Rn) Y

MULTILINEAR DUALITY AND FACTORISATION 49 and d p 1, and thus q 1. Therefore Theorem 1.3 applies. j=1 j ≥ ≥ ExceptP in some rather trivial cases19 we do not know any such explicit factorisations with the 2 sharp constant 1. For example, let v1, v2 and v3 be unit vectors in R with angle 2π/3 between each pair. Then, with Bj being orthogonal projection onto the span of vj , we have 2 (B∗B + B∗B + B∗B )= I . 3 1 1 2 2 3 3 2 2 2 Take qj = 1 for each j so that q = 2. Consequently, for all G L (R ), there exist g1,g2,g3 such that ∈ G(x) g (x)1/3g (x)1/3g (x)1/3 a.e. ≤ 1 2 3 and, for each j, ⊥ ess sups gj (svj + tvj )dt G 2. R ≤k k Z Even in such simple cases as this the factorisation is not yet understood explicitly. 10.2. General Brascamp–Lieb inequalities. On the other hand, under the conditions d (60) pj dim imBj = n j=1 X and d (61) dim V p dimB V ≤ j j j=1 X for all V in the lattice of subspaces of Rn generated by ker B d , we now indicate how to { j }j=1 construct semi-explicit factorisations yielding the finiteness of the constant C in (4). We use the term “semi-explicit” because the construction is algorithmic in nature. Notwithstanding, we give an informal discursive treatment rather than a collection of flow-charts. We assume throughout the discussion that the B are nonzero mappings, that is, n = rank(B ) 1 for each j. (If some j j j ≥ Bj = 0 it plays no role in inequality (4), nor in (60) or (61), and it can simply be dropped.) When n = 1 matters quickly reduce to consideration of H¨older’s inequality, which is treated in Section 9.1 above, so we shall focus on what happens when n 2. ≥ We now sketch how this is done, and we begin with a couple of definitions from [8] and [9]. Given a collection of linear surjections B , its Brascamp–Lieb polytope is defined by { j} d ( B )= (p ,...,p ) [0, )d : dim V p dimB V for all subspaces V . P { j} { 1 d ∈ ∞ ≤ j j } j=1 X This is manifestly a closed convex set, and, as has been previously noted, is contained in [0, 1]d, and is therefore the convex hull of its extreme points. Given data Bj and pj , a critical subspace is a nontrivial proper subspace V of Rn for which { } { } d (62) dim V = pj dimBj V. j=1 X The construction of the factorisations hinges on the question of existence or non-existence of critical subspaces for the problem with data Bj,pj . Indeed, if there is a critical subspace V for B ,p , then the problem of factorising a{ function} on Rn decomposes into two factorisation { j j } subproblems on the spaces V and V ⊥, each of which has positive but strictly smaller dimension

19 For example when the Vj are mutually orthogonal and pj = 1 for all j. 50 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON than n.20 This will allow us in effect to induct on the parameter n. These subproblems inherit the same pj and have “new” Bj which are related to the “old” Bj in a precise way. The two subproblems{ } inherit the conditions corresponding to (60) and (61): indeed, (60) for each of the two subproblems holds precisely because the subspace V is critical, and we shall make crucial use of this fact. We isolate the details of how this works – in particular how factorisations for the two subproblems combine to give a factorisation for the original problem – in Section 10.2.1 below.

On the other hand, if there is no critical subspace for the problem Bj,pj , then (p1,...,pd) lies in the interior of ( B ). To establish a factorisation for the problem{ in} this case, it therefore P { j} suffices to (i) establish factorisations for the extreme points of ( Bj ) and (ii) to show, given factorisations at the extreme points, how to establish factorisatP {ions} at all interior points of ( Bj ). Point (ii) is tantamount to showing that factorisations behave well under multilinear interpolation,P { } and this we have already successfully addressed separately in Section 7.

To deal with point (i), we consider the Brascamp–Lieb problems at the extreme points (˜p1,..., p˜d) 21 of ( Bj ), and, at each of them, ask the same question – does there exist a critical subspace? P { } n Since (˜p1,..., p˜d) is an extreme point of ( Bj ), there will certainly be subspaces V of R satisfying P { }

d (63) dim V = p˜j dimBj V. j=1 X If there is a nontrivial and proper such subspace, we have a critical subspace for the problem Bj, p˜j , and we can proceed as above, in effect going around the loop. The only remaining {possibility} is that the only subspaces V of Rn satisfying (63) are 0 and Rn itself. { } We are thus left to deal with the special case of our original problem in which (p1,...,pd) n is an extreme point of ( Bj ), but for which the only subspaces of V of R satisfying (62) are 0 and Rn itself. MattersP { } quickly reduce to rather trivial considerations. Indeed, in this { } d situation, ( Bj ) consists precisely of those (p1,...,pd) [0, ) lying on the hyperplane d P { } ∈ ∞ j=1 pj nj = n, and its extreme points are precisely those of the form (0,..., 0,n/nj, 0,..., 0). Since n n always, and since ( B ) [0, 1]d, the only circumstances in which this case P j ≤ P { j} ⊆ arises is when nj = n for all j. In this case, our Brascamp–Lieb problem at an extreme point of ( B ) is necessarily of the form (modulo permutations of the coordinate axes) P { j} 1 0 0 1 0 0 f1(B1x) f2(B2x) ...fd(Bdx) dx C f1 f2 ... fd Rn ≤ Rn Rn Rn Z Z  Z  Z  or, equivalently,

f1(B1x)dx C f1 Rn ≤ Rn Z Z  −1 where B1 is invertible. This of course holds with equality with C = (det B1) , and a trivial factorisation applies. Running the machine described above in reverse will thus eventually furnish a factorisation in the general case, and, indeed, the only possible loss in terms of sharp constants occurs at steps where interpolation is employed.

20To facilitate the discussion which follows we should strictly speaking replace the roles of Rn and Rnj by those of abstract n- and nj -dimensional real Hilbert spaces respectively. 21An algorithm for locating these extreme points can be found in [42]. MULTILINEAR DUALITY AND FACTORISATION 51

10.2.1. Factorisation in the presence of a critical subspace. We give the details needed to close the argument set out above in the presence of a critical subspace. The only place we use criticality is that it implies that (60) and (61) hold for the two subproblems which arise – see [8]. Since these are the necessary and sufficient conditions for finiteness of the constant, we may assume that factorisations for the two subproblems exist. (Formally we proceed by induction on n, and the case n = 1 is trivial.) Let B : Rn Rnj be linear surjections. Suppose that U is a nontrivial proper subspace of Rn. j → (As indicated above, we do not assume that it is a critical subspace.) Define B˜ : U B U and j → j B˜ : U ⊥ (B U)⊥ by j → j B˜j (x)= Bj x and ˜ ˜ ⊥ Bj (y) = Π(Bj U) Bj y. ˜ If some B˜j or B˜j is zero we can simply discard it. (It cannot be the case that every B˜j is zero, for d if this happened, we would have U j=1 ker Bj , and, as we have noted previously, a necessary ⊆ ∩ d condition for finiteness of the Brascamp–Lieb constant is that j=1 ker Bj = 0 . For similar ˜ ∩ { } reasons it cannot be the case that every B˜j is zero.) Also define Γ : U ⊥ B U by j → j Γj (y) = Π(Bj U)Bjy. Here, Π denotes orthogonal projection onto a subspace W . So for x U and y U ⊥, W ∈ ∈ B (x + y)= B˜ x + B˜ y +Γ y = B˜ x +Γ y + B˜ y B U (B U)⊥. j j j j j j j ∈ j ⊕ j   The two Brascamp–Lieb subproblems arising can be written in the form

d d pj /p f (B˜ x)pj /pdx C f j j ≤ j U j=1 j=1 Z Y Y Z  and d d pj /p ˜ pj /p fj (B˜j x) dx C fj ⊥ ≤ U j=1 j=1 Z Y Y Z  where p = p 1. With α = p /p, we are entitled to suppose that the following two j j ≥ j j corresponding factorisation statments hold: P ′ For all H Lp (U) of norm 1 there exist H ,...,H such that ∈ 1 d d H(x) H (x)αj ≤ j j=1 Y and, for all φ L1(B U) of norm at most 1, ∈ j φ(B˜ x)H (x)dx K ; j j ≤ 1 ZU ′ For all M Lp (U ⊥) of norm 1 there exist M ,...,M such that ∈ 1 d d M(y) M (y)αj ≤ j j=1 Y 52 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON and, for all ψ L1((B U)⊥) of norm at most 1, ∈ j ˜ ψ(B˜j y)Mj(y)dy K2. ⊥ ≤ ZU ′ Given G Lp (Rn) of norm 1, we want to subfactorise it as ∈ d G(z) G (z)αj ≤ j j=1 Y such that for all f L1(Rnj ) of norm at most 1, ∈

f(Bj z)Gj (z)dz K1K2. Rn ≤ Z This is a factorisation statement corresponding to the problem

d d pj /p pj /p fj(Bj x) dx C fj . Rn ≤ j=1 j=1 Z Y Y Z  If we can do this, then the procedure described above for factorising Brascamp–Lieb problems closes. ′ We begin by writing G Lp of norm 1 as ∈ (64) G(x, y)= Hy(x)M(y) where Hy p′ = 1 for all y and M p′ = 1. We will then factorise M and each Hy as above, and combinek thek factorisations to obtaink k a suitable factorisation for G.

Indeed, defining Hy and M by 1/p′ G(x, y) p′ G(x, y)= ′ G(x, y) dx := Hy(x)M(y) p′ 1/p G(x, y) dx Z  22 is essentially the unique wayR to achieve (64) with the desired conditions. Therefore, d d G(x, y) [H (x)M (y)]αj := G (x, y)αj ≤ jy j j j=1 j=1 Y Y where for all y U ⊥, for all φ L1(B U) of norm at most 1, ∈ ∈ j φ(B˜ x)H (x)dx K j jy ≤ 1 ZU and where for all ψ L1((B U)⊥) of norm at most 1, ∈ j ˜ ψ(B˜j y)Mj(y)dy K2. ⊥ ≤ ZU We want to show that for all f L1(Rnj ) of norm at most 1, ∈

f(Bjz)Gj (z)dz = f(Bj (x, y))Hjy (x)Mj (y)dxdy K1K2. Rn ⊥ ≤ Z ZU ZU 22 r s Indeed, suppose G is in the mixed-norm space Ldy(Ldx) and we want to write G(x,y) = H(x,y)M(y) where s s s kMkr = kGkLr (Ls) and where kH(·,y)ks = 1 for all y. Integrating G(x,y) = H(x,y) M(y) with respect to x shows that the only way to do this is to take M(y) = kG(·,y)ks and H(x,y) = G(x,y)/kG(·,y)ks. See the remarks at the end of Section 9.2. MULTILINEAR DUALITY AND FACTORISATION 53

Fix y U ⊥ and write the inner integral over U as ∈

f(Bj(x, y))Hjy (x)dx = f(Bj x + Bj y)Hjy(x)dx. ZU ZU ˜ ⊥ Now f(Bj x+Bj y)= f((B˜j x+Γj y)+B˜j y). For w BjU and ξ (Bj U) let φξ(w) := f(w +ξ). ˜ ∈ ˜∈ Therefore f(Bj x + Bjy) = φ ˜ (Bj x +Γj y) = (τ(Γj y)φ ˜ )(Bj x), where (τηχ)( ) = χ( + η) Bj y Bj y · · denotes translation by η. So,

˜ f(Bj (x, y))Hjy(x)dx = (τ(Γj y)φ ˜ )(Bj x)Hjy (x)dx K1 τ(Γj y)φ ˜ 1 Bj y ≤ k Bj yk ZU ZU by what we are assuming.

Now, by translation invariance, τ(Γj y)φ ˜ 1 = φ ˜ 1. Therefore, letting ψ(ξ) := φξ 1 for k Bj yk k Bj yk k k ξ (B U)⊥, ∈ j ˜ f(Bj z)Gj(z)dz K1 ψ(B˜j y)Mj(y)dy K1K2 ψ 1. Rn ≤ ⊥ ≤ k k Z ZU Finally,

ψ 1 = φξ(w)dw dξ = f(z)dz =1, k k ⊥ Rnj Z(Bj U) ZBj U ! Z and this gives what we wanted.

11. Multilinear Kakeya inequalities revisited Recall that we have families of 1-tubes in Rn, and for P , its direction e(P ) Sn−1 Pj ∈ Pj ∈ satisfies e(P ) ej cn where cn is a small dimensional constant. The multilinear Kakeya theorem| of Guth− [29|] ≤ (see also [22]) states that

1/n 1/n n n

aPj χPj (x) Cn aPj .   Ln/(n−1)(Rn) ≤   j=1 Pj ∈Pj j=1 Pj ∈Pj Y X Y X   n   This inequality is of the form (7) with X = R , q = n/(n 1), Yj = j with counting measure, p = 1 for all j, α =1/n for all j, and T ((a ))(x)= − a χ P(x). j j Pj Pj ∈Pj Pj Pj Guth proved this result essentially by establishing a suitableP subfactorisation for each nonnegative M Ln(Rn), and then applying Proposition 1.1. His subfactorisation is described in terms of ∈ an auxiliary polynomial p of ‘low’ degree dominated by M n, whose zero set Zp has ‘large’ visibility on each unit cube Q of Rn in the sense that vis(Zk kQ) & M. We do not enter into p ∩ Q the details of the definition of visibility, nor into how this gives the desired subfactorisation, but R instead refer the reader to [29] and [22]. (In the latter paper the approach using Proposition 1.1 is explicit while in the former it is implicit. And one should note that the definition of visibility used in [22] is a power of the original one used in [29].) It was the shock of seeing such an unlikely functional-analytic method succeed which inspired us to study the general question of necessity of factorisation as taken up in this paper. In hindsight, our linkage of subfactorisation of functions with Maurey’s theory of factorisation of operators helps place Guth’s method in perspective. Bourgain and Guth in [16] established an affine invariant form of the multilinear Kakeya inequal- ity, removing the hypothesis that e(P ) e c for P , at the price of inserting a damping | − j |≤ n ∈Pj 54 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON factor on the left-hand side which is consistent with the affine-invariant Loomis–Whitney in- equality of Section 9.2. That is, they proved that for arbitrary families of 1-tubes, Pj 1/(n−1)

aP1 χP1 (x) aPn χPn (x)e(P1) e(Pn) dx Rn ··· ∧···∧ ! Z PX1∈P1 PnX∈Pn 1/(n−1) n C a . ≤ n  Pj  j=1 Y PXj ∈Pj   As the reader will readily verify (using the same argument as in the proof of Proposition 1.1, see also Section 8 above), in order to establish this, it suffices to show that for every nonnegative M Ln(Rn) which is constant on unit cubes in a standard lattice , there exist nonnegative functions∈ S : R such that Q j Q×Pj → S (Q, P )1/n S (Q, P )1/n M(Q) . 1 1 ··· n n e(P ) e(P )1/n 1 ∧···∧ n whenever the 1-tubes P meet at Q, and, for all j, for all P , j j ∈Pj 1/n S (Q, P ) . M(Q)n . j j   Q∈Q,X Q∩Pj 6=∅ QX∈Q And indeed this is what Bourgain and Guth essentially did (see also [22]). It is therefore very tempting to ask whether, in analogy with the situation of Theorem 1.3, this method is guar- anteed to work in so far as the statement of the affine-invariant multilinear Kakeya inequality automatically implies the existence of a subfactorisation as in the last two displayed inequalities. Unfortunately, as we have established above in Section 8, there is no such general functional- analytic principle which guarantees this.

The recent multilinear Kakeya kj -plane inequalities, and indeed the even more general perturbed Brascamp–Lieb inequalities, both recently established by Zhang [45], also fit into the framework we consider, the latter as a generalisation of inequality (5). 11.1. The finite field multilinear Kakeya inequality. Zhang [46] has recently solved the discrete analogue of the multilinear Kakeya problem. Let F be a field and let j be arbitrary families of lines in Fn. For l declare e(l ) e(l ) to be 1 if the vectorsL e(l ) are j ∈ Lj 1 ∧···∧ n { j } linearly independent and to be 0 otherwise. Zhang has proved that for a certain Cn depending only on n, 1/(n−1)

al1 χl1 (x) aln χln (x)e(l1) e(ln) Fn ··· ∧···∧ ! xX∈ l1X∈L1 lnX∈Ln (65) 1/(n−1) n C a . ≤ n  lj  j=1 Y ljX∈Lj   When n = 2 the constant C = 1, as is readily verified using 1/(n 1) = 1 and changing the 2 − order of summation on the left-hand side. Moreover, for general n, if all the lines in j are n L parallel to some fixed vector yj with yj j=1 linearly independent, the constant is likewise 1, since matters can then be reduced to the{ } classical Loomis–Whitney inequality via an invertible linear transformation of Fn, (or one can write down a suitable factorisation as in Example 9.2). MULTILINEAR DUALITY AND FACTORISATION 55

The presence of the factor e(l ) e(l ) in (65) precludes any assertion that (65) is equivalent 1 ∧···∧ n to a factorisation statement: see Section 8 above. If however the j are presumed to satisfy the property that if (l ,...,l ) , then the directionsL e(l ),...,e(l ) are linearly 1 n ∈ L1 ×···×Ln { 1 n } independent, we have that the term e(l1) e(ln) is identically 1, and the result then falls under the scope of Theorem 2.2. ∧···∧ 2 In particular, when n = 2 and we have two finite families of lines 1 and 2 in F such that no line in is parallel to any line in , this holds. Hence we obtain:L L L1 L2 2 Proposition 11.1. Let 1 and 2 be finite families of lines in F such that no line in 1 is L L 2 L parallel to any line in 2. Let J F be the set of points where some l1 1 meets an l2 2. Suppose G(x)2 =1L . Then⊆ there exist g ,g : J R such that for∈ all L x J ∈ L x∈J 1 2 → + ∈ P G(x)= g1(x)g2(x) and, moreover, for all lj j , j =1, 2, p ∈ L g (x) 1. j ≤ x∈XJ∩lj In spite of the extreme simplicity of the original problem, no procedure for coming to an explicit such factorisation is currently known.

Jon Bennett had asked whether, even in higher dimensions, the constant Cn in the finite field F3 multilinear Kakeya inequality might still be 1. This is true in the case of 2. However, this turns out to have been over-optimistic, and we have: Proposition 11.2. Suppose the discrete multilinear Kakeya inequality (65) holds in the case n =3 for F = F3. Then C3 > 1.04.

We remark that Tidor, Yu and Zhao [40] have very recently established numerical values for the constants in Zhang’s theorem, and in particular they show that C √6. 3 ≤ Proof. We construct an example. In this example, for each j =1, 2, 3, we nominate two directions, and the family j will consist of all lines with one of these directions. The two directions for each j will be chosenL so that each of the eight choices of one direction from each of the three families results in a linearly independent set of directions, so that the terms e(l1) e(l2) e(l3) are all 1. Each family of coefficients a – as a function defined on and more properly∧ denoted∧ Lj by aj – is defined to be supported on three lines from j in such a way that the x-summand on the left-hand side of (65) is non-zero at five points. EachL a will take nonzero values in 1, 2 and thus each line under consideration will have a weight equal to 1 or 2. For each j we shall{ } have that two of the three lines pass through two of these five points and the remaining line passes through the remaining point. More concretely, let be the lines with direction (1, 1, 0) or (2, 1, 1); • L1 2 be the lines with direction (0, 1, 0) or (0, 1, 1); and • L be the lines with direction (1, 0, 1) or (0, 0, 1). • L3 It is straightforward to verify that the directions of any three lines, one from each collection, F3 span 3.

We now proceed to properly define the coefficients a. We denote by aj the function whose domain is , and which is defined as follows: Lj Let a1 be • – 2 on the line with direction (1, 1, 0) passing through (0, 2, 2) and (2, 1, 2), 56 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON

– 2 on the line with direction (2, 1, 1) passing through (0, 2, 1) and (2, 0, 2), – 1 on the line with direction (1, 1, 0) through (0, 0, 0), and – 0 on other lines of . L1 Let a2 be • – 2 on the line with direction (0, 1, 0) passing through (2, 0, 2) and (2, 1, 2), – 2 on the line with direction (0, 1, 1) passing through (0, 0, 0) and (0, 2, 2), – 1 on the line with direction (0, 1, 0) through (0, 2, 1), and – 0 on other lines of . L2 Let a3 be • – 2 on the line with direction (0, 0, 1) passing through (0, 2, 1) and (0, 2, 2), – 2 on the line with direction (1, 0, 1) passing through (0, 0, 0) and (2, 0, 2), – 1 on the line with direction (0, 0, 1) through (2, 1, 2), and – 0 for other lines of of . L3 Each has two lines of a-value or weight 2 and one of weight 1. Lj We can see that the only points where lines from all three families intersect are the five points mentioned, namely (0, 0, 0), (0, 2, 1), (0, 2, 2), (2, 0, 2) and (2, 1, 2). At the three points (0, 0, 0), (0, 2, 1) and (2, 1, 2) we have two lines of weight 2 and one of weight 1 meeting; at the two points (0, 2, 2) and (2, 0, 2) we have three lines of weight 2 meeting. So the value of the x-summand on the left-hand side of (65) is 2 at the three points (0, 0, 0), (0, 2, 1) and (2, 1, 2), and is 23/2 at the two points (0, 2, 2) and (2, 0, 2). The left-hand side adds up to 3 2+2 23/2 > 11.65. The · · 3/2 value of the right-hand side of (65) is C 53/2 C 11.19. This shows that C 6+2 > 3 · ≤ 3 · 3 ≥ 53/2 11.65/11.19 > 1.04 > 1. 

A counterexample to the conjecture that (65) holds with Cn = 1 was first found by use of the duality theory developed above, which, as we have mentioned, is valid under the assumption that any n-tuple of lines taken from 1 n has linearly independent directions. To explain why this route was taken, let usL assume×···×L that we are considering a finite field of size q. If we let j consist of all lines with directions in some given set of size r then the input to (65), namely L n−1 the tuple (a1,...,an) belongs to a real vector space of dimension nq r. The input to problem (17) is the function G which belongs to a real vector space of dimension qn. In our case we have n = q = 3 and r = 2 so the input to the problem (17) belongs to a smaller vector space than the input to (65). The additional cost of solving the convex optimisation problem compared with the cost of simply evaluating each side of (65) does not significantly alter the balance of cost. The solution to the convex optimisation problem was found using the software package CVXOPT [1], which yields the solution for both the primal and dual problems. The solution to the dual problem was then slightly simplified by hand for neater exposition and this is what is presented here.

References

[1] M. S. Andersen, J. Dahl, and L. Vandenberghe, CVXOPT: A python package for convex optimization, available at abel.ee.ucla.edu/cvxopt, 2011. [2] J-P. Aubin and I. Ekeland, Applied nonlinear analysis, Pure and Applied Mathematics (New York), John Wiley & Sons Inc., New York, 1984. [3] U. Bader, T. Gelander and N. Monod, A fixed point theorem for L1 spaces, Invent. Math. 189, 1:143–148, 2012. [4] K. M. Ball, Volumes of sections of cubes and related problems, Geometric aspects of (1987–88), (J. Lindenstrauss and V. D. Milman editors), Lecture Notes in Math., 1376, Springer, Berlin:251– 260, 1989. MULTILINEAR DUALITY AND FACTORISATION 57

[5] F. Barthe, On a reverse form of the Brascamp–Lieb inequality, Invent. Math. 134, 2:335–361, 1998. [6] W. Beckner, Inequalities in Fourier Analysis, Ann. of Math. (2), 102, 1:159–182, 1975. [7] J. Bennett, N. Bez, S. Buschenhenke, M. Cowling and T. C. Flock, On the nonlinear Brascamp-Lieb inequality, arXiv:1811.11052 [math.CA], to appear, Duke Math. J. [8] J. Bennett, A. Carbery, M. Christ and T. Tao, The Brascamp–Lieb inequalities: finiteness, structure and extremals, Geom. Funct. Anal. 17, 5:1343–1415, 2008. [9] J. Bennett, A. Carbery, M. Christ and T. Tao, Finite bounds for H¨older–Brascamp–Lieb multilinear inequal- ities, Math. Res. Lett. 17, 4:647–666, 2010. [10] J. Bennett, A. Carbery and T. Tao, On the multilinear restriction and Kakeya conjectures, Acta Math. 196, 2:261–302, 2006. [11] J. Bennett, A. Carbery and J. Wright, A non-linear generalisation of the Loomis–Whitney inequality and applications, Math. Res. Lett. 12, 4:443–457, 2005. [12] J. Bourgain, Moment inequalities for trigonometric polynomials with spectrum in curved hypersurfaces, Israel J. Math. 193 1:441–458, 2013. [13] J. Bourgain, On the Schr¨odinger maximal function in higher dimension, Tr. Mat. Inst. Steklova 280 (2013), Ortogonal’nye Ryady, Teoriya Priblizhenii i Smezhnye Voprosy, 53–66 ISBN: 5-7846-0125-3; 978-5-7846-0125-4; translation in Proc. Steklov Inst. Math. 280 1:46–60, 2013. [14] J. Bourgain and C. Demeter, The proof of the l2 decoupling conjecture, Ann. of Math. (2) 182, 1:351–389, 2015. [15] J. Bourgain, C. Demeter and L. Guth, Proof of the main conjecture in Vinogradov’s mean value theorem for degrees higher than three, Ann. of Math. (2) 184 2:633–682, 2016 [16] J. Bourgain and L. Guth, Bounds on oscillatory integral operators based on multilinear estimates, Geom. Funct. Anal. 21, 6:1239–1295, 2011. [17] S. Boyd and L. Vandenberghe, Convex optimization, Cambridge University Press, Cambridge, 2004. [18] H. J. Brascamp and E. H. Lieb, Best constants in Young’s inequality, its converse, and its generalization to more than three functions, Advances in Math. 20, 2:151–173, (1976). [19] A. P. Calder´on, Intermediate spaces and interpolation, the complex method, Studia Math. 24:113–190, 1964. [20] A. Carbery, T. S. H¨anninen and S. I. Valdimarsson, Disentanglement, multilinear duality and factorisation for non-positive operators, arXiv:2003.03326 [math.FA] [21] A. Carbery, T. S. H¨anninen and S. I. Valdimarsson, manuscript in preparation. [22] A. Carbery and S. I. Valdimarsson, The endpoint multilinear Kakeya theorem via the Borsuk–Ulam theorem. J. Funct. Anal., 264(7):1643–1663, 2013. [23] G. Diestel, Factoring multi-sublinear maps, Jour. Funct. Anal. 266:1928–1947, 2014. [24] N. Dunford and J. Schwartz, Linear Operators, Part 1: General Theory, Wiley Classics Library, 1988. [25] H. Finner, A generalization of H¨older’s inequality and some probability inequalities, Ann. Probab. 20, 4:1893–1901, 1992. [26] J. Garc´ıa-Cuerva and J. L. Rubio de Francia, Weighted Norm Inequalities and Related Topics, North–Holland Mathematics Mathematical Studies 116, North–Holland, Amsterdam, 1985. [27] J. E. Gilbert, Nikisin–Stein Theory and Factorization with applications, in Proc. Symp. Pure Math. XXXV, Part 2:233–267, Amer. Math. Soc., 1979. [28] L. Grafakos and R. H. Torres, A multilinear Schur test and multiplier operators, Jour. Funct. Anal. 187:1–24, 2001. [29] L. Guth, The endpoint case of the Bennett–Carbery–Tao multilinear Kakeya conjecture. Acta Math., 205(2):263–286, 2010. [30] T. H¨anninen and I. Verbitsky, Two-weight Lp → Lq bounds for positive dyadic operators in the case 0 < q ≤ p< ∞, Indiana Univ. Math. J. 69(3):837–871, 2020. [31] E. H. Lieb, Gaussian kernels have only Gaussian maximizers. Invent. Math. 102, 1:179–208, 1990. [32] J. Lindenstrauss and L. Tzafriri, Classical Banach Spaces I and II, Springer Classics in Mathematics, Springer-Verlag Berlin Heidelberg New York, 1977. [33] L. H. Loomis and H. Whitney, An inequality related to the isoperimetric inequality, Bull. Amer. Math. Soc. 55:961–962, 1949. [34] M. Maggis, T. Meyer-Brandis and G. Svindland, Fatou Closedness under Model Uncertainty, Positivity 22(5):1325–1343, 2018. 58 ANTHONYCARBERY,TIMOS.HANNINEN¨ AND STEFAN´ INGI VALDIMARSSON

[35] B. Maurey, Th´eor`emes de factorisation pour les op´eratuers lin´eaires `avaleurs dans les espaces Lp, Ast´erisque 11, Soci´et´eMath´ematique de France, 1974.

[36] G. Pisier, Factorization of Operators through Lp∞ or Lp1 and non-commutative generalizations, Math. Ann., 276(1):105–136, 1986 [37] G. Pisier, Complex interpolation and regular operators between Banach lattices, Arch. Math. 62:261–269, 1994. [38] A. R. Schep, Factorization of positive multilinear maps, Ill. J. Math., 28 (4):579–591, 1984. [39] M. Sch¨onherr and F. Schuricht, Pure Measures, Density Measures and the Dual of L∞, arXiv:1710.02197 [math.MG]. [40] J. Tidor, H. H. Yu and Y. Zhao, Joints of varieties, arXiv:2008.01610 [math.CO] [41] J. F. Toland, The Dual of L∞, Finitely Additive Measures and Weak Convergence, ISBN 978-3-030-34731-4, SpringerBriefs in Mathematics, 2020 [42] S. I. Valdimarsson, The Brascamp-Lieb polyhedron, Canad. J. Math. 62, 4:870–888, 2010. [43] K. Yosida and E. Hewitt, Finitely additive measures, Trans. Amer. Math. Soc., 72:46–66, 1952. [44] A. C. Zaanen, An introduction to the theory of integration, North-Holland Publishing Company, Amsterdam, 1958. [45] R. Zhang, The endpoint perturbed Brascamp-Lieb inequalities with examples, Anal. PDE 11, 3: 555–581, 2018. [46] R. Zhang, A proof of the Multijoints Conjecture and Carbery’s generalization, J. Eur. Math. Soc. 22 (8): 2405–2417, 2020.

Anthony Carbery, School of Mathematics and Maxwell Institute for Mathematical Sciences, Uni- versity of Edinburgh, James Clerk Maxwell Building, Peter Guthrie Tait Road, King’s Buildings, Mayfield Road, Edinburgh, EH9 3FD, Scotland. Email address: [email protected]

Timo S. Hanninen,¨ Department of Mathematics and Statistics, University of Helsinki, P.O. Box 68, FI-00014 Helsinki, Finland, and School of Mathematics and Maxwell Institute for Mathematical Sciences, University of Edinburgh, James Clerk Maxwell Building, Peter Guthrie Tait Road, King’s Buildings, Mayfield Road, Edinburgh, EH9 3FD, Scotland. Email address: [email protected]

Stefan´ Ingi Valdimarsson, Arion banki, Borgartun´ 19, 105 Reykjav´ık, Iceland, and Science Institute, University of Iceland, Dunhagi 5, 107 Reykjav´ık, Iceland Email address: [email protected]