arXiv:1711.10595v3 [math.FA] 24 Oct 2018 christened rtedeksieult,adas ooti etrbound better obtain to also and inequality, Grothendieck’s (i.e., fui -om n h aiu ntergti ae vrall over taken is right the on maximum the and 2-norm, unit of h au fGohnic’ osatdpnso h hieo choice the on depends constant Grothendieck’s of value The where hs xc au ean nnw.Tefloigaesm ma some are following The unknown. remains value exact whose eutta ecle tefnaetltermi h metric the constant in finite a theorem M exists fundamental there “the that called showed he that result 1 max (1) ii it 2]otie lgtysalrbound smaller slightly a obtained [21] Rietz (iii) i)Tepwro rtedeksieult a o general not was inequality Grothendieck’s of power The (ii) i h eta euto rtedeksoiia ae [9] paper original Grothendieck’s of result central The (i) vrtels 5yas hr aebe ayatmt oimpro to attempts many been have there years, 65 last the Over ewl let will We 2010 e od n phrases. and words Key ( = with h infnto o h elcs n bandtecmlxc complex the obtained and case real the argument. for complexification function sign the optn xettoso infntosaduigTyo e for Taylor using bounds and better functions get sign of expectations computing esrusadP cynk 1]1 er ae,wihconn which later, years p 15 Pe lczy´nski [16] and denstrauss a orsodn otesg function. sign the to argume corresponding variational a map using and measure Gaussian normalized ε M smigoeaos hyeuiae n mrvdGrothend improved and elucidated They operators. -summing i h· NEEETR N NFE RO FGROTHENDIECK’S OF PROOF UNIFIED AND ELEMENTARY AN ahmtc ujc Classification. Subject oiu flna ler sprerequisites. as algebra univariate only linear need of that modicum proof a kno a into Haa streamlining Haagerup and and and combining Krivine Krivine, Grothendieck pedagogical, the complex be and both to real yields intended the for and bounds cases explicit known complex and real the Abstract. , = ij ·i ) rtedeksinequality Grothendieck’s π/ ± stesadr ne rdc in product inner standard the is ∈ 1, 2 F F m δ ≤ j × = n epeeta lmnay efcnandpofo Grothend of proof self-contained elementary, an present We = K k , x G R HULFIDAD E-EGLM N IJEZHANG JINJIE AND LIM, LEK-HENG FRIEDLAND, SHMUEL i R ± k = over 1 ≤ rtedekieult,Gohnic osat Krivine constant, Grothendieck inequality, Grothendieck or k y j k sinh( C =1 K G hogotti ril.I 93 rtedekpoe pow a proved Grothendieck 1953, In article. this throughout

R R X i . =1 m π/ ; ε 2) X j i =1 n = n h mletpsil constant possible smallest the and ≈ M 70,4B8 68,8P0 14,0D5 73,4N0 90C 47N10, 97K30, 03D15, 81P45, 81P40, 46B85, 46B28, 47A07, e 1. ij iθ 2 INEQUALITY . i h 0 n 1 and 301 , x > K Introduction F i δ y , j l h aiu ntelf stknoe all over taken is left the on maximum the , j = i uhta o every for that such 0

1 e ≤ iφ K K j . G R 273 over | ≤ ε i | max = 2 ≈ . | nieso Lindenstrauss–Pe lczy´nski, of ideas wn δ 6 n17 yaeaigover averaging by 1974 in 261 C acls ai ope aibe,and variables, complex basic calculus, osat epciey hsatceis article This respectively. constants j sta i pnmu nqaiyholds inequality eponymous his that is | o h rtedekconstant Grothendieck the for s 4 .Teieult 1 a ic been since has (1) inequality The ). =1 /π yrcgie ni h oko Lin- of work the until recognized ly f

eu ons h urn best- current the bounds, gerup F hoyo esrpout”[] he [9]; products” tensor of theory X ce h nqaiyt absolutely to inequality the ected i =1 m o milestones: jor pnin,atog hyddnot did they although xpansions, tt eemn notmlscalar optimal an determine to nt ≤ n ewl eoei by it denote will we and ε i eksieult htunifies that inequality ieck’s δ , X j ekspofi h elcs by case real the in proof ieck’s K =1 n eadsmlf h ro of proof the simplify and ve ,m n m, l, s rmtera aevaa via case real the from ase j G C K M ∈ on,Haeu bound. Haagerup bound, rtedekrle on relied Grothendieck . ij F rtedeksconstant Grothendieck’s ε fui bouevalue absolute unit of i ∈ δ j N

n vr every and x i R y , K n j G F 27. erful with ∈ . K G F F . l , 2 SHMUELFRIEDLAND,LEK-HENGLIM,ANDJINJIEZHANG

R C (iv) Our current best known upper bounds for KG and KG are due to Krivine [14], who in 1979 used theory and ideas in [16] to get R π K 1.78221; G ≤ 2 log(1 + √2) ≈ and Haagerup [10], who in 1987 extended Krivine’s ideas to C to get

C 8 KG 1.40491, ≤ π(x0 + 1) ≈ where x [0, 1] is the unique solution to: 0 ∈ π/2 cos2 t π x dt = (x + 1). 2 2 8 Z0 1 x sin t − R C (v) Our current best known lower boundsp for KG and KG are due to Davie [4, 5], who in 1984 used spherical integrals to get

R 1 ρ(x) KG sup − 1.67696, ≥ x (0,1) max(ρ(x),f(x)) ≈ ∈ where

2 x2/2 2 x2 8 ∞ t2/2 ρ(x) := xe− , f(x) := e− + ρ(x) 1 e− dt ; π π − π r  r Zx  and C 1 θ(x) KG sup − 1.33807, ≥ x>0 g(x) ≈ where 1 x2 ∞ t2 θ(x) := 1 e− + x e− dt , 2 −  Zx  2 1 x2 ∞ t2 2 x2 g(x) := (1 e− )+ e− dt + θ(x) 1 (1 e− ) . x − − x −  Zx    (vi) Progress on improving the aforementioned bounds halted for many years. Believing that R Krivine’s bound is the exact value of KG, some were spurred to find matrices that yield it as R the lower bound of KG [13]. The belief was dispelled in 2011 in a landmark paper [3], which R demonstrated the existence of a positive constant ε such that KG < π/ 2 log(1 + √2) ε but the authors did not provide an explicit better bound. To date, Krivine’s and Haagerup’s− R C  bounds remain the best known explicit upper bounds for KG and KG respectively. (vii) There have also been many alternate proofs of Grothendieck’s inequality employing a variety of techniques, among them factorization of Hilbert spaces [18, 11, 19], absolutely summing operators [7, 16, 20], geometry of Banach spaces [1, 17], metric theory of tensor product [6], basic probability theory [2], bilinear forms on C∗-algebra [12]. In this article, we will present a proof of Grothendieck’s inequality that unifies both the (a) real and (b) complex cases; and yields both the (c) Krivine and (d) Haagerup bounds [14, 10]. It is also elementary in that it requires little more than standard college mathematics. Our proof will rely on Lemma 2.1, which is a variation of known ideas in [16, 10, 11]. In particular, the idea of using the sign function to establish (1) in the real case was due to Grothendieck himself [9] and later also appeared in [16, 14]; whereas the use of the sign function in the complex case first appeared in [10]. To be clear, all the key ideas in our proof were originally due to Lindenstrauss–Pe lczy´nski, Krivine, and Haagerup [16, 14, 10], our only contribution is pedagogical — combining, simplifying, and streamlining their ideas into what we feel is a more palatable proof. To understand the proof, readers need only know univariate calculus, basic complex variables, and a small amount of linear AN ELEMENTARY AND UNIFIED PROOF OF GROTHENDIECK’S INEQUALITY 3 algebra. We will use some basic theory and tensor product constructions in Section 4 but both notions will be explained in a self-contained and elementary way.

2. Gaussian integral of sign function Throughout this article, our inner product over C will be sesquilinear in the second argument, i.e., n x,y := y∗x for all x,y C . h i ∈ For z F = R or C, the sign function is ∈ z/ z if z = 0, (2) sgn(z)= | | 6 (0 if z = 0; and for z Fn, the Gaussian function is ∈ n/2 2 F (2π) exp z /2 if F = R, G (z)= − −k k2 n π n exp( z 2) if F = C. ( − −k k2  Lemma 2.1 below is based on [11, 10]; the complex version in particular is a slight variation of [10, Lemma 3.2]. It plays an important role in our proof because the right side of (3) depends only on the inner product u, v and not (explicitly) on the dimension n. In addition, the functions on the right are homeomorphismsh i and admit Taylor expansions, making it possible to expand them in powers u, v d, which will come in useful when we prove Theorem 4.1. h i Lemma 2.1. Let u, v Fn with u = v = 1. Then ∈ k k2 k k2 2 arcsin u, v if F = R, F π h i (3) sgn u, z sgn z, v Gn(z) dz =  π/2 2 Fn h i h i cos t  u, v dt if F = C. Z h i (1 u, v 2 sin2 t)1/2 Z0 − |h i| Proof. Case I: F = R. Let arccos u, v =θ, so that θ [0,π] and arcsin u, v = π/2 θ. Choose α, β such that 0 < β α<π and defineh i ∈ h i − − E(α, β)= (r cos θ, r sin θ,x3,...,xn) : r 0, α θ β . { ≥ R≤ ≤ } The Gaussian measure of a measurable set A is the integral of Gn (x) over A. Upon integrating with respect to x3,...,xn, the following term remains: β 1 1 (x2+x2) 1 ∞ 1 r2 e− 2 1 2 dx dx = dθ re− 2 dr = (β α)/2π. 2π 1 2 2π − ZE(α,β) Zα Z0 Hence the Gaussian measure of E(α, β) is (β α)/2π. Since there is an isometry T of Rn such that − T u = e1 and T v = (cos θ, sin θ, 0,..., 0), the left side of (3) may be expressed as R sgn T u, x sgn x, T v Gn (x) dx. Rn h i h i Z The set of x where T u, x > 0 and Tv,x > 0 is E(θ π/2, π/2), which has Gaussian measure (π θ)/2π; ditto forh T u, xi < 0 andh Tv,xi < 0. The set− of x where T u, x < 0 and Tv,x > 0 is E−(π/2,θ + π/2), whichh hasi Gaussianh measurei θ/2π; ditto for T u,h x >i 0 and Tv,xh i< 0. The set of x where T u, x = 0 has zero Gaussian measure. Henceh thei value of thish integrai l is (π θ)/2π + (π θ)h/2π iθ/2π θ/2π = 2 arcsin u, v /π. − − − − h i 2n Case II: F = C. We define vectors α, β R with α2i 1 = Re(ui), α2i = Im(ui), β2i 1 = Re(vi), ∈ − 2n − n β2i = Im(vi), i = 1,...,n. Then α and β are unit vectors in R . For any z = (z1,...,zn) C , we write ∈ x = Re(z ), Im(z ),..., Re(z ), Im(z ) R2n. 1 1 n n ∈  4 SHMUELFRIEDLAND,LEK-HENGLIM,ANDJINJIEZHANG

Then, n n Re( u, z )= Re(u z )= Re(u )Re(z )+Im(u )Im(z ) = α, x = x, α , h i i i i i i i h i h i i=1 i=1 X X  and likewise Re( z, v )= x, β . By a change-of-variables and Case I, we have h i h i C R sgn Re u, z sgn Re z, v Gn (z) dz = sgn x, α sgn x, β G2n(x) dx Cn h i h i R2n h i h i Z Z   2 2 (4) = arcsin α, β = arcsin(Re u, v ). π h i π h i It is easy to verify that for any z C, ∈ 2π 1 iθ iθ (5) sgn(z)= sgn(Re(e− z))e dθ. 4 Z0 By (4), (5), and Fubini’s theorem,

C sgn u, z sgn z, v Gn (z) dz Cn h i h i Z 2π 2π 1 iθ iϕ i(θ+ϕ) C = sgn(Re( e− u, z )) sgn(Re( z, e− v ))e Gn (z) dz dθ dϕ 16 0 0 Cn h i h i Z 2π Z 2π Z 1 iθ iϕ i(θ+ϕ) = arcsin(Re( e− u, e v ))e dθ dϕ 8π h i Z0 Z0 Case II(a): u, v R. The integral above becomes h i∈ 1 2π 2π = arcsin(cos(θ + ϕ) u, v )ei(θ+ϕ) dθ dϕ 8π h i Z0 Z0  1 2π 2π+ϕ = arcsin( u, v cos t)eit dt dϕ 8π h i Z0 Zϕ  1 2π 2π 1 2π (6) = arcsin( u, v cos t)eit dt dϕ = arcsin( u, v cos t)eit dt. 8π h i 4 h i Z0 Z0  Z0 Since arcsin( u, v cos t) is an even function with period 2π, h i 2π arcsin( u, v cos t)sin t dt = 0, h i Z0 the last integral in (6) becomes 1 2π arcsin( u, v cos t) cos t dt, 4 h i Z0 and as arcsin( u, v cos t) cos t is an even function with period π, it becomes h i π/2 π/2 arcsin( u, v cos t) cos t dt = arcsin( u, v sin t)sin t dt, h i h i Z0 Z0 which, upon integrating by parts, becomes π/2 cos2 t (7) u, v dt. h i (1 u, v 2 sin2 t)1/2 Z0 − |h i| AN ELEMENTARY AND UNIFIED PROOF OF GROTHENDIECK’S INEQUALITY 5

Case II(b): u, v / R. This reduces to Case II(a) by setting c C of unit modulus so that c u, v = u, vh andi ∈ cu, v R, then by (7), ∈ h i |h i| h i∈ C C sgn u, z sgn z, v Gn (z) dz = c sgn cu, z sgn z, v Gn (z) dz Cn h i h i Cn h i h i Z Z π/2 cos2 t π/2 cos2 t = c cu, v dt = u, v dt. h i (1 cu, v 2 sin2 t)1/2 h i (1 u, v 2 sin2 t)1/2 Z0 − |h i| Z0 − |h i| 

We will make a simple but useful observation1 about the quantities in (1) that we will need for the proof of Corollary 2.3 later. Lemma 2.2. Let F = R or C and d,m,n N. For any M Fm n, we have ∈ ∈ × m n m n (8) max Mijεiδj = max Mijεiδj εi 1, δj 1 εi = δj =1 | |≤ | |≤ i=1 j=1 | | | | i=1 j=1 X X X X Fd and for any x1,...,xm,y1,...,y n , m ∈n m n (9) max Mij xi,yj = max Mij xi,yj . xi 1, yj 1 h i xi = yj =1 h i k k≤ k k≤ i=1 j=1 k k k k i=1 j=1 X X X X m n Proof. We will start with (8). Suppose there exists M F × such that the left-hand side of (8) ∈ exceeds the right-hand side. Let the maximum of the left-hand side be attained by ε1∗,...,εm∗ and δ1∗,...,δn∗ . By our assumption, at least one εi∗ or δj∗ must be less than 1 in absolute value and so let ε < 1 without loss of generality. Fix ε = ε , i = 2,...,m and δ = δ , j = 1,...,n, but let ε vary | 1∗| i i∗ j j∗ 1 with ε 1 and consider the maximum of the left hand-side over ε . Since max aε +b : ε 1 | 1|≤ 1 {| 1 | | 1|≤ } is always attained on the boundary ε1 = 1 for any a, b F, this contradicts our assumption. The proof for (9) is similar with norm in| place| of absolute value.∈  In the corollary below, the inequality on the left is the “original Grothendieck inequality,” i.e., as first stated by Grothendieck2 in [9], and the inequality on the right is due to Haagerup [10]. Corollary 2.3. Let F = R or C and d,m,n N. For any M Fm n with ∈ ∈ × m n (10) max Mijεiδj 1, εi = δj =1 ≤ | | | | i=1 j=1 X X Fd any x1,...,xm,y1,...,yn of unit 2-norm, we have ∈ m n m n π M arcsin x ,y if F = R, M H( x ,y ) 1 if F = C, ij h i ji ≤ 2 ij h i ji ≤ i=1 j=1 i=1 j=1 X X X X F C where H denotes the function on the right side of (3) for = . Proof. The condition (10) implies that m n R R M sgn x ,x sgn y ,x G (z) G (z), ij h i i h j i d ≤ d i=1 j=1 X X m n C C M sgn z, x sgn z, y G (z) G (z), ij h ii h ji d ≤ d i=1 j=1 X X

1This of course follows from other well-known results but we would like to keep our exposition self-contained. 2The better known modern version (1) is in fact due to Lindenstrauss and Pe lczy´nski in [16]. 6 SHMUELFRIEDLAND,LEK-HENGLIM,ANDJINJIEZHANG for any x Rd, z Cd respectively. Integrating over Rd or Cd respectively and applying Lemma 2.1 give the required∈ results.∈ Note that we have implicitly relied on (8) in Lemma 2.2 as the sgn function is not always of absolute value one and may be zero.  Corollary 2.3 already looks a lot like the Grothendieck inequaltiy (1) but the nonlinear functions arcsin and H are in the way. To obtain the Grothendieck inequality, we linearize them: First by using Taylor series to replace these functions by polynomials; and then using a ‘tensor trick’ to express the polynomials as linear functions on a larger space. This is the gist of the proof in Section 4.

3. Haagerup function We will need to make a few observations regarding the functions on the right side of (3) for the proof of Grothendieck’s inequality. Let the complex Haagerup function of a complex variable z be π/2 cos2 t H(z) := z dt, z 1, (1 z 2 sin2 t)1/2 | |≤ Z0 − | | and the real Haagerup function h as the restriction of H to [ 1, 1] R. Observe that h : [ 1, 1] [ 1, 1] and is a strictly increasing continuous bijection. Since−[ 1, 1]⊆ is compact, h is a homeomor-− → phism− of [ 1, 1] onto itself. By the Taylor expansion − − 2 2 1/2 ∞ (2k 1)!! 2k 2k (1 x sin t)− = − x sin t, x 1, 0 t<π/2, − (2k)!! | |≤ ≤ Xk=0 and π/2 π (2k 1)!! cos2 t sin2k t dt = − , 4(k + 1) (2k)!! Z0 thus we get 2 ∞ π (2k 1)!! (11) h(x)= − x2k+1, x [ 1, 1]. 4(k + 1) (2k)!! ∈ − Xk=0   1 Since h is analytic at x = 0 and h′(0) = 0, its inverse function h− : [ 1, 1] [ 1, 1] can be expanded in a power series in some neighborhood6 of 0 − → −

1 ∞ 2k+1 (12) h− (x)= b2k+1x . Xk=0 One may in principle determine the coefficients using the Lagrange inversion formula: 1 d2k t 2k+1 b2k+1 = lim . (2k + 1)! t 0 dt2k h(t) →     For example, 4 1 4 3 1 4 7 b = , b = , b = 0, b = . 1 π 3 −8 π 5 7 −1024 π But determining b2k+1 explicitly becomes difficult as k gets larger. A key step  in Haagerup’s proof [10] requires the nonpositivity of the coefficients beyond the first: (13) b 0, for all k 1. 2k+1 ≤ ≥ This step is in our view the most technical part of [10]. We have no insights on how it may be avoided but we simplified Haagerup’s proof of (13) in Section 5 to keep to our promise of an elementary proof — using only calculus and basic complex variables. It follows from (13) that h(z) := b z h 1(z) has nonnnegative Taylor coefficients. Pringsheim’s 1 − − theorem implies that if the radius of convergence of the Taylor series of h(z) is r, then h(z), and e e e AN ELEMENTARY AND UNIFIED PROOF OF GROTHENDIECK’S INEQUALITY 7

1 thus h− (z), has a singular point at z = r. As h′(t) > 0 on (0, 1) and h(1) = 1, we must have 1 N 2k+1 N r 1. It also follows from (13) that h− (t) k=0 b2k+1t for any t (0, 1) and N . So ≥N 2k+1 1 ≤ N N ∈ ∈ k=1 b2k+1 t b1t h− (t) for any t (0, 1) and N . So k=1 b2k+1 b1 1 for any N| | ≤ − ∈ P1 ∈ | | ≤ − 1 N and we have k∞=0 b2k+1 2b1 1. As h− (1) = 1 we deduce that k∞=0 b2k+1 = h− (1) = 1, andP ∈ therefore | |≤ − P P P ∞ (14) b = 2b 1. | 2k+1| 1 − Xk=0 We now turn our attention back to the complex Haagerup function. Observe that H(z) = h( z ) for all z D := z C : z 1 and arg(H(z)) = arg(z) for 0 = z D. So H| : D| D| |is ∈ { ∈ | | ≤ } 1 6 ∈ → a homeomorphism of D onto itself. Let H− : D D be its inverse function. Since H(z) = sgn(z)h( z ), we get → | | 1 1 ∞ 2k+1 (15) H− (z) = sgn(z)h− ( z ) = sgn(z) b z . | | 2k+1| | Xk=0 2k+1 Dini’s theorem shows that the function ϕ(x) := ∞ b x is a strictly increasing and k=0 | 2k+1| continuous on [0, 1], with ϕ(0) = 0 and ϕ(1) = ∞ b2k+1 b1 = 4/π > 1; note that ϕ(1) is Pk=0 | | ≥ finite by (14). Thus there exists a unique c0 (0, 1) such that ϕ(c0) = 1. So ∈ P ∞ 2k+1 8 1 1= ϕ(c )= b c = c h− (c ), 0 | 2k+1| 0 π 0 − 0 Xk=0 1 where the last equality follows from b1 = 4/π and (13). Therefore we obtain h− (c0) = 8c0/π 1, and if we let x := h 1(c ) (0, 1), then h(x ) π(x + 1)/8 = 0. From the Taylor expansion− of 0 − 0 ∈ 0 − 0 h(x), the function x h(x) π(x + 1)/8 is increasing and continuous on [0, 1]. Hence x0 is the unique solution in [0,7→1] to − π (16) h(x) (x + 1) = 0 − 8 and c0 = π(x0 + 1)/8. As Corollary 2.3 indicates, the Haagerup function H plays the analogue of arcsin in the complex case. Unlike arcsin, H is a completely obscure function,3 and any of its properties that we require will have to be established from scratch. The goal of this section is essentially to establish (11)–(16), which we will need later.

4. A unified proof of Grothendieck’s inequality In this section we will need the notions of (i) tensor product and (ii) Hilbert space, but just F Fn (2k+1) F R C enough to make sense of n( )= k∞=0( )⊗ where = or . In keeping to our promise of an elementary proof, weH will briefly introduce these notions in a simple manner. For our purpose, it suffices to regard the tensor productL of k copies of Fn, denoted n k n n (F )⊗ = F F , ⊗···⊗ k copies F as the -vector space of k-dimensional hypermatrices| {z, } Fn k F ( )⊗ := [ai1 i ] : ai1 i , i1,...,ik 1,...,n , ··· k ··· k ∈ ∈ { } where scalar multiplication and vector addition of hypermatrices are defined coordinatewise. For k vectors x,y,...,z Fn, their tensor product is the k-dimensional hypermatrix given by ∈ n n k x y z := [x y z ] (F )⊗ . ⊗ ⊗···⊗ i1 i2 · · · ik i1,i2,...,ik=1 ∈ 3We are unaware of any other occurrence of H outside its use in Haagerup’s proof of his bound in [10]. 8 SHMUELFRIEDLAND,LEK-HENGLIM,ANDJINJIEZHANG

We write k x⊗ := x x . ⊗···⊗ k copies If , is an inner product on Fn, then defining h· ·i | {z } (17) x y z,x′ y′ z′ := x,x′ y,y′ z, z′ h ⊗ ⊗···⊗ ⊗ ⊗···⊗ i h ih i···h i n k and extending bilinearly (if F = R) or sesquilinearly (if F = C) to all of (F )⊗ yields an inner product on the k-dimensional hypermatrices. In particular we have k k k x⊗ ,y⊗ = x,y . h i h i If e ,...,e is the standard orthonormal basis of Fn, then { 1 n} n k (18) e e (F )⊗ : i ,...,i 1,...,n i1 ⊗···⊗ ik ∈ 1 k ∈ { } n k is an orthonormal basis of (F )⊗ . For more information about hypermatrices see [15] and for a more formal definition of tensor products see [8]. If an F-vector space is equipped with an inner product , such that every Cauchy sequence H h· ·i in converges with respect to the induced norm v = v, v 1/2, we call a Hilbert space. Hilbert spacesH need not be finite-dimensional; we call separablek k |h ifi| there is a countableH set of orthonormal vectors e : j J , i.e., J is a countableH index set, such that every v satisfies { j ∈H ∈ } ∈H (19) v 2 = v, e 2. k k |h j i| j J X∈ n (2k+1) Let , k be the inner product on (F )⊗ as defined in (17), k be its induced norm, and B beh· the·i orthonormal basis in (18). Let n N. The F-vector spacek4 · k k ∈ ∞ n (2k+1) n (2k+1) ∞ 2 (20) n(F) := (F )⊗ = (v0, v1, v2,... ) : vk (F )⊗ , vk < H ∈ k=0k kk ∞ Mk=0 n X o equipped with the inner product

∞ (21) u, v := uk, vk k h i∗ h i Xk=0 is a separable Hilbert space since k∞=0 Bk is a countable set of orthonormal vectors satisfying (19). We write for the norm induced by (21). k · k∗ S Theorem 4.1 (Grothendieck inequality with Krivine and Haagerup bounds). Let F = R or C and l,m,n N. For any M Fm n, any x ,...,x ,y ,...,y Fl of unit 2-norm, we have ∈ ∈ × 1 m 1 n ∈ m n m n F (22) max Mij xi,yj K max Mijεiδj , xi = yj =1 h i ≤ εi = δj =1 k k k k i=1 j=1 | | | | i=1 j=1 X X X X where R π C 8 K := and K := 2 log(1 + √2) π(x0 + 1) are Krivine’s and Haagerup’s bounds respectively. Recall that x0 is as defined in (16). Proof. As we described at the end of Section 2, we will ‘linearize’ the nonlinear functions arcsin and H in Corollary 2.3 by using Taylor series to replace these functions by polynomials, followed by a ‘tensor trick’ to express polynomials as linear functions on an infinite-dimensional space.

4The direct sum in (20) is a Hilbert space direct sum, i.e., it is the closure of the vector space direct sum. AN ELEMENTARY AND UNIFIED PROOF OF GROTHENDIECK’S INEQUALITY 9

Case I: F = R. Let c := arcsinh(1) = log(1 + √2). Taylor expansion gives

2k+1 2k+1 ∞ k c 2k+1 ∞ k c (2k+1) (2k+1) (23) sin(c x ,y )= ( 1) x ,y = ( 1) x⊗ ,y⊗ . h i ji − (2k + 1)!h i ji − (2k + 1)! i j k k=0 k=0 X X For any l N, let (R) be as in (20), and S, T : Rl (R) be nonlinear maps defined by ∈ Hl →Hl

2k+1 k c (2k+1) S(x) := Sk(x) ∞ , Sk(x) = ( 1) x⊗ , k=0 − s(2k + 1)! ·  2k+1 c (2k+1) T (x) := Tk(x) ∞ , Tk(x)= x⊗ , k=0 s(2k + 1)! ·  l for any x R . To justify that S and T are indeed maps into l(R), we need to demonstrate that S(x) , ∈T (x) < but this follows from H k k∗ k k∗ ∞

2k+1 2 ∞ 2 ∞ c 2(2k+1) ∞ 2 2 S(x) = Sk(x) k = x = Tk(x) k = T (x) k k∗ k k (2k + 1)!k k k k k k∗ Xk=0 Xk=0 Xk=0 and

∞ c2k+1 x 2(2k+1) = sinh(c x 2) < (2k + 1)!k k k k ∞ Xk=0 for all x Rl. Note that ∈

∞ c2k+1 S(x), T (y) = ( 1)k x,y 2k+1 = sin(c x,y ). h i∗ − (2k + 1)!h i h i Xk=0 Hence (23) becomes:

sin(c xi,yj )= S(xi), T (yj) or c xi,yj = arcsin S(xi), T (yj) . h i h i∗ h i h i∗

l Moreover, since xi and yj are unit vectors in R , we get

S(x ) 2 = sinh(c x 2)=1 and T (y ) 2 = sinh(c y 2) = 1. k i k k ik k j k k j k

As the m + n vectors S(x1),...,S(xm), T (y1),...,T (yn) in l(R) span a subspace l(R) of dimension d m + n; and since any two finite-dimensional innerH product spaces areS⊆H isometric, is isometric to≤ Rd with the standard inner product. So we may apply Corollary 2.3 to obtain S

m n m n 1 π Mij xi,yj = Mij arcsin S(xi), T (yj) , h i c h i∗ ≤ 2c i=1 j=1 i=1 j=1 X X X X

which is Krivine’s bound since π/2c = π/ 2 log(1 + √2) = KR.  10 SHMUELFRIEDLAND,LEK-HENGLIM,ANDJINJIEZHANG

Case II: F = C. Let c0 (0, 1) be the unique constant defined in (16) such that ϕ(c0) = 1. By the Taylor expansion in (15)∈ and noting that sgn(z) z 2k+1 = zkzk+1, | | 1 ∞ 2k+1 H− (c x ,y ) = sgn(c x ,y ) b c x ,y 0h i ji 0h i ji 2k+1| 0h i ji| Xk=0 ∞ k = b c2k+1 x ,y x ,y k+1 2k+1 0 h i ji h i ji Xk=0 ∞ = b c2k+1 x , y k x ,y k+1 2k+1 0 h i ji h i ji Xk=0 ∞ 2k+1 k (k+1) k (k+1) (24) = b c x⊗ x⊗ , y⊗ y⊗ . 2k+1 0 i ⊗ i j ⊗ j k k=0 X l For any l N, let Dl = x C : x 1 be the unit ball, let l(C) be as in (20), and let S, T : D ∈ (C) be nonlinear{ ∈ maps definedk k ≤ by} H l →Hl 2k+1 1/2 (k) (k+1) S(x)= S (x) ∞ , S (x) := sgn(b ) b c x¯⊗ x⊗ , k k=0 k 2k+1 | 2k+1| 0 · ⊗ 2k+1 1/2 (k) (k+1) T (x)= T (x)∞ , T (x) := b c x¯⊗  x⊗ , k k=0 k | 2k+1| 0 · ⊗ for any x D . Then S and T are maps into (C) since ∈ l  Hl  2 ∞ 2 ∞ 2k+1 2(2k+1) ∞ 2 2 S(x) = Sk(x) k = b2k+1 c0 x = Tk(x) k = T (x) k k∗ k k | | k k k k k k∗ Xk=0 Xk=0 Xk=0 and, as b > 0 and b 0 for all k 1 by (13), 1 2k+1 ≤ ≥ ∞ 2k+1 2(2k+1) 2 1 2 b c x = 2b c x H− (c x ) < . | 2k+1| 0 k k 1 0k k − 0k k ∞ Xk=0 As in Case I, we may rewrite (24) as 1 H− (c0 xi,yj )= S(xi), T (yj) or c0 xi,yj = H( S(xi), T (yj) ). h i h i∗ h i h i∗ l Moreover, since xi and yj are unit vectors in C , we get

∞ S(x ) 2 = b c2k+1 = ϕ(c ) = 1, k i k | 2k+1| 0 0 Xk=0 and similarly T (y ) = 1. So we may apply Corollary 2.3 to get k j k m n m n 1 1 Mij xi,yj = MijH( S(xi), T (yj) ) , h i c h i∗ ≤ c i=1 j=1 0 i=1 j=1 0 X X X X C  which is Haagerup’s bound since 1/c 0 = 8/π (x0 +1) = K .

5. Nonpositivity of b2k+1 To make the proof in this article entirely self-contained, we present Haagerup’s proof of the nonpositivity of b2k+1 that we used earlier in (13). While the main ideas are all due to Haagerup, our small contribution here is that we avoided the use of any known results of elliptic integrals in order to stay faithful to our claim of an elementary proof, i.e., one that uses only calculus and basic complex variables. To be clear, while the functions π/2 π/2 2 2 1/2 2 2 1/2 (25) K(x) := (1 x sin t)− dt, E(x) := (1 x sin t) dt − − Z0 Z0 AN ELEMENTARY AND UNIFIED PROOF OF GROTHENDIECK’S INEQUALITY 11 do make a brief appearance in the proof of Lemma 5.1, the reader does not need to know that they are the complete elliptic integrals of the first and second kinds respectively. Haagerup had relied liberally on properties of K and E that require substantial effort to establish [10]. We will only use trivialities that follow immediately from definitions. Our point of departure from Haagerup’s proof is the following lemma about two functions h1 and h2, which we will see in Lemma 5.2 arise respectively from the real and imaginary parts of the analytic extension of the real Haagerup function h : [ 1, 1] [ 1, 1] to the upper half plane. − → − Lemma 5.1. Let h , h : [1, ) R be defined by 1 2 ∞ → π/2 h (x) := 1 x 2 sin2 t dt, 1 − − Z0 p π/2 2 2 sin t h (x) := (1 x− ) dt, 2 2 − 0 1 (1 x 2)sin t Z − − − which are clearly strictly increasing functions on p[1, ) with ∞ h1(1) = 1, lim h1(x)= π/2, h2(1) = 0, lim h2(x)= . x x →∞ →∞ ∞ Then π (26) ω (x) := x(h (x)h′ (x) h′ (x)h (x)) = for x 1, 1 1 2 − 1 2 2 ≥ π (27) ω (x) := x(h (x)h′ (x)+ h (x)h′ (x)) 2h (√2)h (√2) > for 1 x √2. 2 1 1 2 2 ≥ 1 2 4 ≤ ≤

Proof. We start by observing some properties of h1′ and h2′ . As 1 π/2 sin2 t 1 π/2 sin2 t h1′ (x)= 3 dt = 2 dt, x 0 1 x 2 sin2 t x 0 x2 sin2 t Z − − Z − π/2 p 1 +p h1′ is strictly decreasing on (1, ). As 0 cos− t dt = , limx 1 h1′ (x)= . Clearly limx h1′ (x)= ∞ ∞ → ∞ →∞ 0. Furthermore, when x> 1, since x2 sin2 t √x2 1, we have R − ≥ − p π (28) 0 < h1′ (x) for x> 1. ≤ 4x2√x2 1 − It is straightforward to see that the functions E and K in (25) have derivatives given by

1 1 2 (29) E′(y)= E(y) K(y) , K′(y)= E(y) (1 y )K(y)) . y − y(1 y2) − − −   Clearly, h (x)= K(y) E(y), where y = y(x)= √1 x 2. So by chain rule, 2 − − − π/2 d 1 2 2 1/2 h′ (x)= y′(x) (K E)(y(x)) = (1 (1 x− )sin t) dt. 2 dy − x − − Z0 Hence h2′ is strictly decreasing on [1, ), h2′ (1) = π/2, and limx h2′ (x) = 0. To show (26), observe that ∞ →∞

h (x)= E(1/x), xh′ (x)= K(1/x) E(1/x), h (x)= K(y) E(y), xh′ (x)= E(y), 1 1 − 2 − 2 where again y = √1 x 2. Hence − − ω (x)= E(1/x)E(y) [K(1/x) E(1/x)][K(y) E(y)] 1 − − − = E(1/x)K(y)+ K(1/x)E(y) K(1/x)K(y). − Computing ω1′ , we see from (29) that ω1′ 0. So ω1 is a constant function. By (28), limx 1 h1′ (x)(1 2 ≡ ց − x− ) = 0, and so limx 1 ω1(x)= π/2. Thus ω1(x)= π/2 for all x> 1 and we may set ω1(1) = π/2. ց 12 SHMUELFRIEDLAND,LEK-HENGLIM,ANDJINJIEZHANG

We now show (27) following Haagerup’s arguments. Note that ω (x)= E(1/x)(K(1/x) E(1/x)) + E(y)(K(y) E(y)). 2 − − Let g(x) := E(√x)(K(√x) E(√x)). A straightforward calculation using (29) shows that − 1 E(√x) K(√x) E(√x) 2 g′′(x)= − 0, x [0, 1]. 2 1 x − x ≥ ∈  −  So g is convex on [0, 1]. Hence g(1 x) is also convex on [0, 1]. Let f(x) := g(x)+ g(1 x). Then f is convex on [0, 1] and f (1/2) =− 0. Therefore f(x) f(1/2) 2g(1/2). This yields− the first ′ ≥ ≥ inequality in (27): ω (x) 2h (√2)h (√2) for x [1, √2]. 2 ≥ 1 2 ∈ The Taylor expansions of h1 and h2 may be obtained as that in (11), 2 π ∞ (2k)! 1 2k (30) h (x)= x− , 1 2 22k(k!)2 1 2k Xk=0  − 2 π ∞ (2k)! 2k 2 k (31) h (x)= (1 x− ) . 2 2 22k(k!)2 2k 1 − Xk=0  − 5 Approximate numerical values of h1 and h2 at x = √2 and 4 are calculated to be: (32) h (√2) 1.3506438, h (√2) 0.5034307, h (4) 1.5459572, h (4) 1.7289033. 1 ≈ 2 ≈ 1 ≈ 2 ≈ The second inequality in (27) then follows from 2h (√2)h (√2) 2 1.35064 0.50343 > π/4.  1 2 ≈ × × In the next two lemmas and their proofs, Arg will denote principal argument. Lemma 5.2. Let h : [ 1, 1] [ 1, 1] be the real Haagerup function as defined in Section 3. Then − → − h can be extended to a function h+ : H C that is continuous on the closed upper half-plane H = z C : Im(z) 0 and analytic on→ the upper half-plane H = z C : Im(z) > 0 . In { ∈ ≥ } { ∈ } addition, h+ has the following properties:

(i) Im(h+(z)) Im(h+( z )) for all z H z C : z 1 and h+(z) = 0 for all z H 0 . (ii) For x [1,≥ ), | | ∈ ∩ { ∈ | |≥ } 6 ∈ \{ } ∈ ∞ Re(h+(x)) = h1(x), Im(h+(x)) = h2(x),

where h1, h2 are as defined in Lemma 5.1. (iii) For all k N and all real α> 1, ∈ α 2 (2k+1) (33) b = Im (h (x))− dx + r (α) 2k+1 π(2k + 1) + k Z1 where  α (2k+1) (34) r (α) Im(h (α)) − . | k |≤ 2k + 1 + Proof. Integrating by parts, we obtain  π/2 π/2 h(x)= cos t d(arcsin(x sin t)) = sin t arcsin(x sin t) dt, x [ 1, 1]. · ∈ − Z0 Z0 The analytic function sin z is a bijection of [ π/2, π/2] [0, ) onto H and it maps the line segment − × ∞ t + ia : π/2 t π/2 onto the half ellipsoid z H : z 1 + z + 1 = 2 cosh a . Let arcsin { − ≤ ≤ } { ∈ | − | | | } +

5For example, using http://www.wolframalpha.com, which is freely available. Such numerical calculations cannot be completely avoided — Haagerup’s proof implicitly contains them as he used tabulated values of elliptic integrals. AN ELEMENTARY AND UNIFIED PROOF OF GROTHENDIECK’S INEQUALITY 13 be the inverse of this mapping. Then arcsin+ is continuous in H and analytic in H. In addition, we have: arcsin x if x [ 1, 1], arcsin+ x = ∈ − π sgn x + i arccosh x if x ( , 1) (1, ), ( 2 | | ∈ −∞ − ∪ ∞ 1 Im(arcsin z) = arccosh ( z 1 + z + 1 ) , z H. + 2 | − | | | ∈   If we define π/2 h (z) := sin t arcsin (z sin t) dt, z H, + + ∈ Z0 then h+ is a continuous extension of h to H and is analytic in H. (i) Since arccosh is increasing on [1, ), we have ∞ 1 arccosh z if z 1, Im(arcsin+ z) = arccosh ( z 1 + z + 1 ) | | | |≥ 2 | − | | | ≥ (0 if z < 1.   | | Therefore for z H z C : z 1 , ∈ ∩ { ∈ | |≥ } π/2 Im(h (z)) = sin t Im(arcsin (z sin t)) dt + · + Z0 π/2 sin t arccosh( z sin t) dt = Im(h+( z )). ≥ arcsin(1/ z ) | | | | Z | | As Im(arcsin+ z) > 0 on H, we have Im(h+(z)) > 0 on H. For x [ 1, 1], h+(x) = h(x) is zero only at x = 0. For x ( , 1) (1, ), ∈ − ∈ −∞ − ∪ ∞ π/2 Im(h+(x)) = sin t arccosh( x sin t) dt > 0. arcsin(1/ x ) | | Z | |

Hence h+ has no zero in H 0 . (ii) Let x (1, ). Integrating\{ by} parts followed by a change-of-variables sin u = x sin t in the next-to-last∈ ∞ equality gives us: arcsin(1/x) π π/2 Re(h (x)) = sin t arcsin(x sin t) dt + sin t dt + 2 Z0 Zarcsin(1/x) arcsin(1/x) 2 π/2 cos t 2 2 = x dt = 1 x− sin u du = h1(x). 0 1 x2 sin2 t 0 − Z − Z p A change-of-variables sin v = (1p x 2) 1/2 cos t in the next-to-last equality gives us: − − − π/2 π/2 cos2 t Im(h+(x)) = sin t arccosh(x sin t) dt = x dt 2 2 Zarcsin(1/x) Zarcsin(1/x) x sin t 1 π/2 2 − 2 sin v p = (1 x− ) dv = h (x). 2 2 − 0 1 (1 x 2)sin v Z − − − (iii) The power series (11) shows that hpdefines an analytic function h(z) in the open unit disk that is identically equal to h+(z) on z C : z < 1 H. Since h(0) = 0 and h′(0) = 0, we can find some δ (0, 1] such that h({z)∈ has an| analytic| }∩ inverse function (12) in z C6 : z < δ . 0 ∈ { ∈ | | 0} For 0 <δ<δ0, let Cδ be a counterclockwise orientated circle with radius δ. It follows 14 SHMUELFRIEDLAND,LEK-HENGLIM,ANDJINJIEZHANG

that h(Cδ) is a simple closed curve with winding number +1. Integrating by parts with a change-of-variables, we have 1 1 h− (z) 1 z b = dz = h′(z) dz. 2k+1 2πi z2k+2 2πi h(z)2k+2 Zh(Cδ) ZCδ Note that b R and 2k+1 ∈ zh (z) 1 d z (2k + 1) ′ dz + dz = dz = 0. − h(z)2k+2 h(z)2k+1 dz h(z)2k+1 ZCδ ZCδ ZCδ   Then we get

1 (2k+1) 1 (2k+1) b = h(z)− dz = Im(h(z)− ) dz 2k+1 2π(2k + 1) 2π(2k + 1) ZCδ ZCδ 2 (2k+1) = Im(h(z)− ) dz π(2k + 1) C′ Z δ where C is the quarter circle δeiθ : 0 θ π/2 . Since h(z) identically equals h (z) on C δ′ { ≤ ≤ } + δ′ and h+(z) has no zeros in the set z C : δ z α, 0 Arg z π/2 by (i), Cauchy’s integral formula yields { ∈ ≤ | | ≤ ≤ ≤ } α iδ 2 (2k+1) (2k+1) (2k+1) b2k+1 = Im h+(z)− dz + h+(z)− dz + h+(z)− dz . π(2k + 1) ′ Zδ ZCα Ziα  Moreover, since h+(z) is real on [δ, 1] and its real part vanishes on the imaginary axis, we are left with α 2 (2k+1) 2 (2k+1) b2k+1 = Im(h+(z)− ) dz + Im h+(z)− dz . π(2k + 1) π(2k + 1) ′ Z1 ZCα  By (i), h (z) Im(h (z)) Im(h ( z )). Thus + ≥ + ≥ + | | (2k+1) πα (2k+1) h+(z)− dz Im(h+(α)) − .  ′ ≤ 2 ZCα  The integral expression of b in (33) will be an important ingredient in the proof that b 0 2k+1 2k+1 for k 1. We establish some further approximations for this integral in the next and final lemma.≤ ≥ Lemma 5.3. Let α = 4 throughout.6 Let θ(x) := Arg(h (x)) for x [1, ). Then θ : [1, ) + ∈ ∞ ∞ → [0, 2π] is strictly increasing on for x 1, θ(1) = 0, and limx θ(x)= π/2. In addition, we have the following: ≥ →∞ (i) Let p := (2k + 1)θ(α)/π . Let ⌊ ⌋ θ(x)=πr/(2k+1) 2 (2k+1) Ir := h+(x) − sin((2k + 1)θ(x)) dx π(2k + 1) θ(x)=π(r 1)/(2k+1) | | | | Z − for r = 1, 2,...,p, and α 2 (2k+1) J := h (x) − sin((2k + 1)θ(x)) dx. π(2k + 1) | + | | | Zθ(x)=πp/(2k+1) Then α 2 (2k+1) p p+1 Im(h (x)− ) dx = I + I . . . + ( 1) I + ( 1) J. π(2k + 1) + − 1 2 − − p − Z1 (ii) Let k 4. Then p 2 and I >I > >I > J. ≥ ≥ 1 2 · · · p 6To avoid confusion, we write ‘α’ for the upper limit of our integrals instead of ‘4’ as the same number will also appear in an unrelated context ‘k ≥ 4.’ AN ELEMENTARY AND UNIFIED PROOF OF GROTHENDIECK’S INEQUALITY 15

(iii) Let k 4 and c = h (√2) e θ(√2)/2. Then I > 0.57c (2k+1)/(2k + 1)2 and I < 0.85I . ≥ | + | − 1 − 2 1 Proof. Since θ(x) = arctan(h2(x)/h1(x)), by (26), we get dθ(x) h (x)h (x) h (x)h (x) (35) = 1 2′ − 1′ 2 > 0, x> 1. dx h (x) 2 | + | So θ(x) is strictly increasing on for x 1. It is clear that θ(1) = 0. By Lemma 5.1, limx h1(x)= ≥ →∞ π/2 and limx h2(x)=+ , so limx θ(x)= π/2. →∞ ∞ →∞ (i) This follows from dividing the interval of the integral [1, α] into p + 1 subsets: α α 2 (2k+1) 2 (2k+1) Im(h (x)− ) dx = h (x) − sin((2k + 1)θ(x)) dx π(2k + 1) + −π(2k + 1) | + | Z1 Z1 = I + I . . . + ( 1)pI + ( 1)p+1J. − 1 2 − − p − (ii) We write x = x(θ), θ [0, π/2), for the inverse function of θ = θ(x). By (35), we have ∈ πr/(2k+1) 4 2k+1 Ir = 2 x(θ) h+(x(θ)) − sin((2k + 1)θ) dθ, π (2k + 1) π(r 1)/(2k+1) | | | | Z − θ(α) 4 2k+1 J = x(θ) h (x(θ)) − sin((2k + 1)θ) dθ. π2(2k + 1) | + | | | Zπp/(2k+1) By Lemma 5.1, h1(x) and h2(x) are strictly increasing function of x [1, ), therefore, so 2 2 2 2k+1∈ ∞ is h+(x) = h1(x) + h2(x) . With this, we deduce that x h+(x) − is strictly decreasing on| [1, α]| for k 4 as | | ≥ d 2k+1 2k+1 ( 2k + 1)x 2k 1 d 2 (x h (x) − )= h (x) − + − h (x) − − h (x) dx | + | | + | 2 | + | dx| + | 2k 1 2 = h (x) − − h (x) (2k 1)x(h (x)h′ (x)+ h (x)h′ (x)) | + | | + | − − 1 1 2 2 2k 1 2 π h (x) − − h (x) (2k 1)  ≤ | + | | + | − − 4 2k 1 2 7π  h (x) − − h (α) 0.1187 < 0, ≤ | + | | + | − 4 ≈− where we have used the fact that h (x) 2 is increasing on [1, α] in the next-to-last inequality | + | and the numerical value is calculated from those of h1(4) and h2(4) in (32). Since sin((2k + 1)θ) is periodic with period π/(2k + 1), we obtain I >I >...>I . In addition,| | 1 2 p θ(α) 4 2k+1 J = x(θ) h (x(θ)) − sin((2k + 1)θ) dθ π2(2k + 1) | + | | | Zπp/(2k+1) θ(α) π/(2k+1) 4 − 2k+1 2 x(θ) h+(x(θ)) − sin((2k + 1)θ) dθ π/4 = arctan(1), and so p = (2k + 1)θ(α)/π 9θ(α)/π = 2 for k 4. ≈ (iii) Since⌊ x(θ) 1 for⌋ ≥θ ⌊ [0, π/2),⌋ we have ≥ ≥ ∈ π/(2k+1) 4 2k+1 I x(θ) h (x(θ)) − sin((2k + 1)θ) dθ. 1 ≥ π2(2k + 1) | + | | | Z0 Recall that θ = θ(x) and x = x(θ) are inverse functions of one another. For θ [0,θ(√2)], ∈ d 1 d dθ 1 log h (x(θ)) = log h (x) 2 − dθ | + | 2 dx | + | · dx h (x)h (x)+ h (x)h (x) ω (x) 1 = 1 1′ 2 2′ = 2 > , h (x)h (x) h (x)h (x) ω (x) 2 1 2′ − 1′ 2 1 16 SHMUELFRIEDLAND,LEK-HENGLIM,ANDJINJIEZHANG

for x [1, √2], where we have used (26), (27), and the fact that θ(x) is strictly increasing for ∈ x 1. Hence log h (x(θ)) log h (√2) (θ(√2) θ)/2 which is equivalent to ≥ | + |≤ | + |− − h (x(θ)) ceθ/2, θ [0,θ(√2)] | + |≤ ∈ θ(√2)/2 where c = h+(√2) e− 1.2059 and θ(√2) > π/9, using values of h1(√2) and h2(√2) in (32). | | ≈ It follows that for k 4, we have ≥ π/(2k+1) 4 θ/2 2k+1 I (ce )− sin((2k + 1)θ) dθ 1 ≥ π2(2k + 1) Z0 2k+1 π 2k+1 π 4c− (k 1/2)θ/(2k+1) 4c− θ/2 = e− − sin θ dθ e− sin θ dθ π2(2k + 1)2 ≥ π2(2k + 1)2 Z0 Z0 2c 2 1+ e π/2 c (2k+1) 0.57c (2k+1) = − − > − . π 1 + 1/4 · (2k + 1)2 (2k + 1)2   Since d log h (x(θ)) 1/2, we get dθ | + |≥ 2k+1 (k 1/2)π/(2k+1) 2k+1 h (x(θ + π/(2k + 1))) − e− − h (x(θ)) − | + | ≤ | + | Moreover, since θ(5/√3) > 2π/9, we know that x(θ) 5/√3 on [0, 2π/9]. Hence for k 4, it follows from the above results that ≤ ≥ 2π/(2k+1) 4 5 2k+1 Ir = x(θ) h+(x(θ)) − sin((2k + 1)θ) dθ π2(2k + 1) √3 | | | | Zπ/(2k+1) π/(2k+1) 4 5 2k+1 = x(θ) h+(x(θ + π/(2k + 1))) − sin((2k + 1)θ) dθ π2(2k + 1) √3 | | Z0 π/(2k+1) 4 5 (k 1/2)π/(2k+1) 2k+1 e− − x(θ) h+(x(θ)) − sin((2k + 1)θ) dθ ≤ π2(2k + 1) √3 | | Z0 5 7π/18 e− I1 < 0.85I1.  ≤ √3

The fact that x h (x) 2k+1 is strictly decreasing on [1, 4] for k 4, established in the proof of | + |− ≥ (ii) above, is a crucial observation for establishing the nonpositivity of b2k+1 for k 4. Observe ≥ 7 that since h+(x) is strictly increasing for x > 1, it is enough to show that x h+(x) − is strictly | | | 2k+1| decreasing on [1, 4], which is what we did. Note that for a fixed k 1, x h+(x) − is increasing for large enough x, as h (x) behaves like C log x for x 1. ≥ | | | + | ≫ Theorem 5.4. Let the Taylor expansion of h 1(x) be as in (12). Then b 0 for k 1. − 2k+1 ≤ ≥ Proof. Let k 4 and let I1,I2,...,Ip, J be as defined in Lemma 5.3. By (33) with α = 4 and Lemma 5.3(i)≥ and (ii), we have

p 1 p b = I I + . . . + ( 1) − I + ( 1) J r (5√2) >I I r (5√2). − 2k+1 1 − 2 − p − − 2k+1 1 − 2 − 2k+1 By (34) and Lemma 5.3(iii) with c 1.2059 (established in its proof), we get ≈ 0.0855 (2k+1) 4 (2k+1) I I > (1.206)− , r (4) (1.728)− . 1 − 2 (2k + 1)2 | 2k+1 |≤ 2k + 1

Since b2k+1 > I1 I2 r2k+1(4), we get b2k+1 < 0 for k 9. Direct computation using the Lagrange− inversion formula− − gives us b , b , . . . , b 0, proving≥ nonpositivity for k 8.  3 5 17 ≤ ≤ AN ELEMENTARY AND UNIFIED PROOF OF GROTHENDIECK’S INEQUALITY 17

Acknowledgment We thank the anonymous referee for his careful reading and many helpful suggestions that greatly improved our article. The work in this article is generously supported by DARPA D15AP00109 and NSF IIS 1546413. LHL acknowledges additional support from a DARPA Director’s Fellowship and the Eckhardt Faculty Fund. References [1] F. Albiac and N. J. Kalton. Topics in Banach space theory, volume 233 of Graduate Texts in Mathematics. Springer, [Cham], second edition, 2016. With a foreword by Gilles Godefory. [2] R. C. Blei. An elementary proof of the Grothendieck inequality. Proc. Amer. Math. Soc., 100(1):58–60, 1987. [3] M. Braverman, K. Makarychev, Y. Makarychev, and A. Naor. The Grothendieck constant is strictly smaller than Krivine’s bound. Forum Math. Pi, 1:e4, 42, 2013. [4] A. M. Davie. Lower bound for kg. Unpublished note, 1984. [5] A. M. Davie. Matrix norms related to Grothendieck’s inequality. In Banach spaces (Columbia, Mo., 1984), volume 1166 of Lecture Notes in Math., pages 22–26. Springer, Berlin, 1985. [6] J. Diestel, J. H. Fourie, and J. Swart. The metric theory of tensor products. American Mathematical Society, Providence, RI, 2008. Grothendieck’s r´esum´erevisited. [7] J. Diestel, H. Jarchow, and A. Tonge. Absolutely summing operators, volume 43 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 1995. [8] S. Friedland and M. Aliabadi. Linear Algebra and Matrices. SIAM, Philadelphia, PA, 2018. [9] A. Grothendieck. R´esum´ede la th´eorie m´etrique des produits tensoriels topologiques. Bol. Soc. Mat. S˜ao Paulo, 8:1–79, 1953. [10] U. Haagerup. A new upper bound for the complex Grothendieck constant. Israel J. Math., 60(2):199–224, 1987. [11] G. J. O. Jameson. Summing and nuclear norms in Banach space theory, volume 8 of London Mathematical Society Student Texts. Cambridge University Press, Cambridge, 1987. [12] S. Kaijser. A simple-minded proof of the Pisier-Grothendieck inequality. In Banach spaces, harmonic analysis, and probability theory, volume 995 of Lecture Notes in Math., pages 33–44. Springer, Berlin, 1983. [13] H. K¨onig. On an extremal problem originating in questions of unconditional convergence. In Recent progress in multivariate approximation (Witten-Bommerholz, 2000), volume 137 of Internat. Ser. Numer. Math., pages 185–192. Birkh¨auser, Basel, 2001. [14] J.-L. Krivine. Constantes de Grothendieck et fonctions de type positif sur les sph`eres. Adv. in Math., 31(1):16–30, 1979. [15] L.-H. Lim. Tensors and hypermatrices Chapter 15, 30 pp., Handbook of Linear Algebra, 2nd Ed., CRC Press, Boca Raton, FL, 2013. [16] J. Lindenstrauss and A. Pe lczy´nski. Absolutely summing operators in Lp-spaces and their applications. Studia Math., 29:275–326, 1968. [17] J. Lindenstrauss and L. Tzafriri. Classical Banach spaces. I. Springer-Verlag, Berlin-New York, 1977. Sequence spaces, Ergebnisse der Mathematik und ihrer Grenzgebiete, Vol. 92. [18] B. Maurey. Une nouvelle d´emonstration d’un th´eor`eme de Grothendieck. page 7, 1973. [19] G. Pisier. Factorization of linear operators and geometry of Banach spaces, volume 60 of CBMS Regional Con- ference Series in Mathematics. Published for the Conference Board of the Mathematical Sciences, Washington, DC; by the American Mathematical Society, Providence, RI, 1986. [20] G. Pisier. Grothendieck’s theorem, past and present. Bull. Amer. Math. Soc. (N.S.), 49(2):237–323, 2012. [21] R. E. Rietz. A proof of the Grothendieck inequality. Israel J. Math., 19:271–276, 1974. [22] J. Zhang, S. Friedland, and L.-H. Lim. Grothendieck constant is norm of strassen matrix multiplication tensor. arXiv:1711.04427.

Department of Mathematics, Statistics and Computer Science, University of Illinois, Chicago, IL 60607-7045 E-mail address: [email protected]

Computational and Applied Mathematics Initiative, Department of Statistics, University of Chicago, Chicago, IL 60637-1514 E-mail address, corresponding author: [email protected]

Department of Statistics, University of Chicago, Chicago, IL 60637-1514 Current address: Department of Mathematics, University of California, San Diego, CA 92093-0112 E-mail address: [email protected]