Quasi-invariant Gaussian measures for the cubic fourth order nonlinear Schrödinger equation

Similar documents
TADAHIRO OH 0, 3 8 (T R), (1.5) The result in [2] is in fact stated for time-periodic functions: 0, 1 3 (T 2 ). (1.4)

ON THE PROBABILISTIC CAUCHY THEORY FOR NONLINEAR DISPERSIVE PDES

The NLS on product spaces and applications

Almost Sure Well-Posedness of Cubic NLS on the Torus Below L 2

The Gaussian free field, Gibbs measures and NLS on planar domains

Global well-posedness for KdV in Sobolev spaces of negative index

A COUNTEREXAMPLE TO AN ENDPOINT BILINEAR STRICHARTZ INEQUALITY TERENCE TAO. t L x (R R2 ) f L 2 x (R2 )

Very Weak Turbulence for Certain Dispersive Equations

Brownian Motion. 1 Definition Brownian Motion Wiener measure... 3

Almost sure global existence and scattering for the one-dimensional NLS

9 Radon-Nikodym theorem and conditioning

Energy transfer model and large periodic boundary value problem for the quintic NLS

Global solutions for the cubic non linear wave equation

Lecture 22 Girsanov s Theorem

SCATTERING FOR THE TWO-DIMENSIONAL NLS WITH EXPONENTIAL NONLINEARITY

Global well-posedness for semi-linear Wave and Schrödinger equations. Slim Ibrahim

DETERMINATION OF THE BLOW-UP RATE FOR THE SEMILINEAR WAVE EQUATION

Citation Osaka Journal of Mathematics. 41(4)

The Dirichlet s P rinciple. In this lecture we discuss an alternative formulation of the Dirichlet problem for the Laplace equation:

CUTOFF RESOLVENT ESTIMATES AND THE SEMILINEAR SCHRÖDINGER EQUATION

Brownian motion. Samy Tindel. Purdue University. Probability Theory 2 - MA 539

Partial Differential Equations

The Schrödinger equation with spatial white noise potential

TADAHIRO OH AND CATHERINE SULEM

ELEMENTS OF PROBABILITY THEORY

THE INVERSE FUNCTION THEOREM

A PEDESTRIAN APPROACH TO THE INVARIANT GIBBS MEASURES FOR THE 2-d DEFOCUSING NONLINEAR SCHRÖDINGER EQUATIONS

1 Math 241A-B Homework Problem List for F2015 and W2016

ON THE REGULARITY OF SAMPLE PATHS OF SUB-ELLIPTIC DIFFUSIONS ON MANIFOLDS

4 Inverse function theorem

Applications of Ito s Formula

GLOBAL WELL-POSEDNESS OF NLS-KDV SYSTEMS FOR PERIODIC FUNCTIONS

RENORMALIZATION OF DYSON S VECTOR-VALUED HIERARCHICAL MODEL AT LOW TEMPERATURES

4 Hilbert spaces. The proof of the Hilbert basis theorem is not mathematics, it is theology. Camille Jordan

at time t, in dimension d. The index i varies in a countable set I. We call configuration the family, denoted generically by Φ: U (x i (t) x j (t))

1 Assignment 1: Nonlinear dynamics (due September

2 A Model, Harmonic Map, Problem

3 Integration and Expectation

Finite Differences for Differential Equations 28 PART II. Finite Difference Methods for Differential Equations

Integration on Measure Spaces

Invariant measures and the soliton resolution conjecture

DISPERSIVE EQUATIONS: A SURVEY

NOTES ON EXISTENCE AND UNIQUENESS THEOREMS FOR ODES

THE FORM SUM AND THE FRIEDRICHS EXTENSION OF SCHRÖDINGER-TYPE OPERATORS ON RIEMANNIAN MANIFOLDS

Erratum to: Logarithm laws for flows on homogeneous spaces

THE INVERSE FUNCTION THEOREM FOR LIPSCHITZ MAPS

1 Lyapunov theory of stability

ANALYSIS & PDE. mathematical sciences publishers TADAHIRO O H PERIODIC STOCHASTIC KORTEWEG DE VRIES EQUATION WITH ADDITIVE SPACE-TIME WHITE NOISE

PROBABILITY: LIMIT THEOREMS II, SPRING HOMEWORK PROBLEMS

LINEAR ALGEBRA W W L CHEN

STAT 7032 Probability Spring Wlodek Bryc

M ath. Res. Lett. 15 (2008), no. 6, c International Press 2008 SCATTERING FOR THE NON-RADIAL 3D CUBIC NONLINEAR SCHRÖDINGER EQUATION

Lecture 9 Metric spaces. The contraction fixed point theorem. The implicit function theorem. The existence of solutions to differenti. equations.

Lecture 5: Hodge theorem

The Sine-Gordon regime of the Landau-Lifshitz equation with a strong easy-plane anisotropy

Lecture No 1 Introduction to Diffusion equations The heat equat

Lecture 9. d N(0, 1). Now we fix n and think of a SRW on [0,1]. We take the k th step at time k n. and our increments are ± 1

Random Bernstein-Markov factors

Regularity of the density for the stochastic heat equation

here, this space is in fact infinite-dimensional, so t σ ess. Exercise Let T B(H) be a self-adjoint operator on an infinitedimensional

Laplace s Equation. Chapter Mean Value Formulas

Lecture Notes 1: Vector spaces

DERIVATION OF THE CUBIC NLS AND GROSS-PITAEVSKII HIERARCHY FROM MANYBODY DYNAMICS IN d = 2, 3 BASED ON SPACETIME NORMS

HILBERT SPACES AND THE RADON-NIKODYM THEOREM. where the bar in the first equation denotes complex conjugation. In either case, for any x V define

Functional Analysis. Franck Sueur Metric spaces Definitions Completeness Compactness Separability...

NOTES ON THE REGULARITY OF QUASICONFORMAL HOMEOMORPHISMS

Kernel Method: Data Analysis with Positive Definite Kernels

On the stability of filament flows and Schrödinger maps

A PHYSICAL SPACE PROOF OF THE BILINEAR STRICHARTZ AND LOCAL SMOOTHING ESTIMATES FOR THE SCHRÖDINGER EQUATION

Piecewise Smooth Solutions to the Burgers-Hilbert Equation

PARTIAL DIFFERENTIAL EQUATIONS. Lecturer: D.M.A. Stuart MT 2007

A GLOBAL COMPACT ATTRACTOR FOR HIGH-DIMENSIONAL DEFOCUSING NON-LINEAR SCHRÖDINGER EQUATIONS WITH POTENTIAL TERENCE TAO

YAN GUO, JUHI JANG, AND NING JIANG

Theory of PDE Homework 2

The continuity method

9 Brownian Motion: Construction

Some SDEs with distributional drift Part I : General calculus. Flandoli, Franco; Russo, Francesco; Wolf, Jochen

Smoothing Effects for Linear Partial Differential Equations

in Bounded Domains Ariane Trescases CMLA, ENS Cachan

Exponential Mixing Properties of Stochastic PDEs Through Asymptotic Coupling

Wiener Measure and Brownian Motion

Periodic Sinks and Observable Chaos

) ) = γ. and P ( X. B(a, b) = Γ(a)Γ(b) Γ(a + b) ; (x + y, ) I J}. Then, (rx) a 1 (ry) b 1 e (x+y)r r 2 dxdy Γ(a)Γ(b) D

Krzysztof Burdzy University of Washington. = X(Y (t)), t 0}

BLOWUP THEORY FOR THE CRITICAL NONLINEAR SCHRÖDINGER EQUATIONS REVISITED

The heat equation in time dependent domains with Neumann boundary conditions

SCATTERING FOR THE NON-RADIAL 3D CUBIC NONLINEAR SCHRÖDINGER EQUATION

Remarks on Bronštein s root theorem

A LOWER BOUND ON BLOWUP RATES FOR THE 3D INCOMPRESSIBLE EULER EQUATION AND A SINGLE EXPONENTIAL BEALE-KATO-MAJDA ESTIMATE. 1.

A G Ramm, Implicit Function Theorem via the DSM, Nonlinear Analysis: Theory, Methods and Appl., 72, N3-4, (2010),

Exponential approach to equilibrium for a stochastic NLS

Notes for Expansions/Series and Differential Equations

Growth of Sobolev norms for the cubic NLS

Solutions to the Nonlinear Schrödinger Equation in Hyperbolic Space

SPECTRAL PROPERTIES OF THE LAPLACIAN ON BOUNDED DOMAINS

THEODORE VORONOV DIFFERENTIABLE MANIFOLDS. Fall Last updated: November 26, (Under construction.)

Scattering for the NLS equation

Conservative Control Systems Described by the Schrödinger Equation

BIHARMONIC WAVE MAPS INTO SPHERES

McGill University Department of Mathematics and Statistics. Ph.D. preliminary examination, PART A. PURE AND APPLIED MATHEMATICS Paper BETA

Transcription:

Probab. Theory Relat. Fields (2017) 169:1121 1168 DOI 10.1007/s00440-016-0748-7 Quasi-invariant Gaussian measures for the cubic fourth order nonlinear Schrödinger equation Tadahiro Oh 1,2 Nikolay Tzvetkov 3 Received: 4 August 2015 / Revised: 25 November 2016 / Published online: 23 December 2016 The Author(s) 2016. This article is published with open access at Springerlink.com Abstract We consider the cubic fourth order nonlinear Schrödinger equation on the circle. In particular, we prove that the mean-zero Gaussian measures on Sobolev spaces H s (T), s > 4 3, are quasi-invariant under the flow. Keywords Fourth order nonlinear Schrödinger equation Biharmonic nonlinear Schrödinger equation Gaussian measure Quasi-invariance Mathematics Subject Classification 35Q55 Contents 1 Introduction... 1122 1.1 Background... 1122 1.2 Cubic fourth order nonlinear Schrödinger equation... 1 1.3 Main result... 1124 1.4 Organization of the paper... 1126 2 Notations... 1126 3 Reformulation of the cubic fourth order NLS... 1127 B Tadahiro Oh hiro.oh@ed.ac.uk Nikolay Tzvetkov nikolay.tzvetkov@u-cergy.fr 1 School of Mathematics, The University of Edinburgh, James Clerk Maxwell Building, The King s Buildings, Peter Guthrie Tait Road, Edinburgh EH9 3FD, UK 2 The Maxwell Institute for the Mathematical Sciences, James Clerk Maxwell Building, The King s Buildings, Peter Guthrie Tait Road, Edinburgh EH9 3FD, UK 3 Université de Cergy-Pontoise, 2, Av. Adolphe Chauvin, 95302 Cergy-Pontoise Cedex, France

1122 T. Oh, N. Tzvetkov 4 Gaussian measures under transformations... 1129 5 Ramer s argument: s > 1... 1133 5.1 Normal form reduction... 1134 5.2 Consequence of Ramer s result... 1136 6 Proof of Theorem 1.2: s > 4 3... 1141 6.1 Energy estimate... 1142 6.2 Weighted Gaussian measures... 1146 6.3 A change-of-variable formula... 1149 6.4 On the evolution of the truncated measures... 1151 6.5 Proof of Theorem 1.2... 1153 Appendix A. On the Cauchy problem (1.1)... 1155 Well-posedness in L 2 (T)... 1155 Ill-posedness below L 2 (T)... 1159 Appendix B. On the approximation property of the truncated dynamics... 1161 References... 1165 1 Introduction 1.1 Background In this paper, we continue the program set up by the second author [70] and study the transport property of Gaussian measures on Sobolev spaces under the dynamics of a certain Hamiltonian partial differential equation (PDE). In probability theory, there is an extensive literature on the transport property of Gaussian measures under linear and nonlinear transformations. See, for example, [3,6,19,24,25,45,62]. Classically, Cameron Martin [19] studied the transport property of Gaussian measures under a shift and established a dichotomy between absolute continuity and singularity of the transported measure. In the context of nonlinear transformations, the work in [45,62] considers nonlinear transformations that are close to the identity, while the work in [24,25] considers the transport property under the flow generated by (non-smooth) vector fields. In particular, in [25], the existence of quasi-invariant measures under the dynamics was established under an exponential integrability assumption of the divergence of the corresponding vector field. We also note a recent work [53] establishing absolute continuity of the Gaussian measure associated to the complex Brownian bridge on the circle under certain gauge transformations. In the field of Hamiltonian PDEs, Gaussian measures naturally appear in the construction of invariant measures associated to conservation laws such as Gibbs measures. These invariant measures associated to conservation laws are typically constructed as weighted Gaussian measures. There has been a significant progress over the recent years in this subject. See [8 12,14 16,18,27,28,30,31,48 50,52,55 57,61,63,65,67 69,71,72,74,75]. On the one hand, in the presence of such an invariant weighted Gaussian measure, one can study the transport property of a specific Gaussian measure, relying on the mutual absolute continuity of the invariant measure and the Gaussian measure. On the other hand, the invariant measures constructed in the forementioned work are mostly supported on rough functions with the exception of completely integrable Hamiltonian PDEs such as the cubic nonlinear Schrödinger equation (NLS), the KdV equation, and the Benjamin Ono equation [29,71,72,74,75].

Quasi-invariant Gaussian measures for the cubic fourth order 1 These completely integrable equations admit conservation laws at high regularities, allowing us to construct weighted Gaussian measures supported on smooth functions. In general, however, it is rare to have a conservation law at a high regularity and thus one needs an alternative method to study the transport property of Gaussian measures supported on smooth functions under the dynamics of non-integrable PDEs. In the following, we consider the cubic fourth order NLS as a model equation and study the transport property of Gaussian measures supported on smooth functions. In particular, we prove that the transported Gaussian measures and the original Gaussian measures are mutually absolutely continuous with respect to each other. Our approach combines PDE techniques such as an energy estimate and normal form reductions and probabilistic techniques in an intricate manner. 1.2 Cubic fourth order nonlinear Schrödinger equation As a model dispersive equation, we consider the cubic fourth order nonlinear Schrödinger equation on T: { i t u = x 4u ± u 2 u (x, t) T R, (1.1) u t=0 = u 0, where u is a complex-valued function on T R with T = R/(2πZ).TheEq.(1.1)is also called the biharmonic NLS and it was studied in [40,66] in the context of stability of solitons in magnetic materials. The biharmonic NLS (1.1) is a special case of the following more general class of fourth order NLS: i t u = λ 2 x u + μ 4 x u ± u 2 u. (1.2) The model (1.2) was introduced in [41,42] to include the effect of small fourth-order dispersion terms in the propagation of intense laser beams in a bulk medium with Kerr nonlinearity. See also [5,33,60] for the references therein. The Eq. (1.1) is a Hamiltonian PDE with the following Hamiltonian: H(u) = 1 x 2 2 u 2 dx ± 1 u 4 dx. (1.3) T 4 T Moreover, the mass M(u) defined by M(u) = u 2 dx (1.4) T is conserved under the dynamics of (1.1). This mass conservation allows us to prove the following global well-posedness of (1.1) inl 2 (T). Proposition 1.1 The cubic fourth order NLS (1.1) is globally well-posed in H s (T) for s 0.

1124 T. Oh, N. Tzvetkov See Appendix A for the proof. We point out that Proposition 1.1 is sharp in the sense that (1.1) is ill-posed below L 2 (T). See the discussion in Sect. A.2. See also [38,59]. Our main goal is to study the transport property of Gaussian measures on Sobolev spaces under the dynamics of (1.1). 1.3 Main result We first introduce a family of mean-zero Gaussian measures on Sobolev spaces. Given s > 1 2,letμ s be the mean-zero Gaussian measure on L 2 (T) with the covariance operator 2(Id ) s, written as dμ s = Z 1 s e 1 2 u 2 H s du = Z 1 s e 1 2 n 2s û n 2 dû n. (1.5) While the expression dμ s = Zs 1 exp( 2 1 u 2 H s )du may suggest that μ s is a Gaussian measure on H s (T), we need to enlarge a space in order to make sense of μ s. The Gaussian measure μ s defined above is in fact the induced probability measure under the map 1 n Z ω u ω (x) = u(x; ω) = n Z g n (ω) n s einx, (1.6) where =(1 + 2 ) 1 2 and {g n } n Z is a sequence of independent standard complexvalued Gaussian random variables, i.e. Var(g n ) = 2. Note that u ω in (1.6) lies in H σ (T) for σ<s 1 2 but not in H s 1 2 (T) almost surely. Moreover, for the same range of σ, μ s is a Gaussian probability measure on H σ (T) and the triplet (H s, H σ,μ s ) forms an abstract Wiener space. See [36,46]. Recall the following definition of quasi-invariant measures. Given a measure space (X,μ), we say that μ is quasi-invariant under a transformation T : X X if the transported measure T μ = μ T 1 and μ are equivalent, i.e. mutually absolutely continuous with respect to each other. We now state our main result. Theorem 1.2 Let s > 3 4. Then, the Gaussian measure μ s is quasi-invariant under the flow of the cubic fourth order NLS (1.1). When s = 2, one may obtain Theorem 1.2 by establishing invariance of the Gibbs measure dρ = Z 1 exp( H(u))du and appealing to the mutual absolute continuity of the Gibbs measure ρ and the Gaussian measure μ 2, at least in the defocusing case. Such invariance, however, is a very rigid statement and is not applicable to other values of s > 3 4. Instead, we follow the approach introduced by the second author in the context of the (generalized) BBM equation [70]. In particular, we combine both PDE techniques and probabilistic techniques in an intricate manner. Moreover, we perform both local and 1 In the following, we drop the harmless factor of 2π.

Quasi-invariant Gaussian measures for the cubic fourth order 1125 global analysis on the phase space. An example of local analysis is an energy estimate (see Proposition 6.1 below), where we study a property of a particular trajectory, while examples of global analysis include the transport property of Gaussian measures under global transformations discussed in Sect. 4 and a change-of-variable formula (Proposition 6.6). As in [70], it is essential to exhibit a smoothing on the nonlinear part of the dynamics of (1.1). Furthermore, we crucially exploit the invariance property of the Gaussian measure μ s under some nonlinear (gauge) transformation. See Sect. 4. In the context of the generalized BBM considered in [70], there was an obvious smoothing coming from the smoothing operator applied to the nonlinearity. There is, however, no apparent smoothing for our Eq. (1.1). In fact, a major novelty compared to [70] is that in this work we exploit the dispersive nature of the equation in a fundamental manner. Our main tool in this context is normal form reductions analogous to the approach employed in [4,37,47]. In [4], Babin Ilyin Titi introduced a normal form approach for constructing solutions to dispersive PDEs. It turned out that this approach has various applications such as establishing unconditional uniqueness [37,47] and exhibiting nonlinear smoothing [32]. The normal form approach is also effective in establishing a good energy estimate, though such an application of the normal form reduction in energy estimates is more classical and precedes the work of [4]. See Sect. 6.1. In [62], Ramer proved a criterion on quasi-invariance of a Gaussian measure on an abstract Wiener space under a nonlinear transformation. In the context of our problem, this result basically states that μ s is quasi-invariant if the nonlinear part is (1 + ε)- smoother than the linear part. See [45] for a related previous result. In Sect. 5, we perform a normal form reduction on the renormalized Eq. (3.6) and exhibit (1 + ε)- smoothing on the nonlinear part if s > 1. This argument provides the first proof of Theorem 1.2 when s > 1. It seems that the regularity restriction s > 1 is optimal for the application of Ramer s result. See Remark 5.4. When s 1, we need to go beyond Ramer s argument. In this case, we follow the basic methodology in [70], combining an energy estimate and global analysis of truncated measures. Due to a lack of apparent smoothing, our energy estimate is more intricate. Indeed, we need to perform a normal form reduction and introduce a modified energy for this purpose. This introduces a further modification to the argument from [70]. See Sect. 6. Lastly, let us point out the following. While the regularity restriction s > 4 3 in Theorem 1.2 comes from the energy estimate (Proposition 6.1), we expect that, by introducing some new ideas related to more refined normal form reductions developed in [37], the result may be extended to the (optimal) regularity range s > 2 1. We plan to address this question in a future work. Remark 1.3 (i) In the higher regularity setting s > 1, we can reduce the proof of Theorem 1.2 to Ramer s result [62]. See Sect. 5. While there is an explicit representation for the Radon Nikodym derivative in [62], we do not know how to gain useful information from it at this point (ii) In the low regularity case 4 3 < s 1, we employ the argument introduced in [70]. See Sect. 6. This argument is more quantitative and in particular, it allows us to obtain a polynomial upper bound on the growth of the Sobolev norm. However, such a polynomial growth bound may also be obtained by purely deterministic methods.

1126 T. Oh, N. Tzvetkov See Remark 7.4 in [70]. A quasi-invariance result with better quantitative bounds may lead to an improvement of the known deterministic bounds. At the present moment, however, we do not know how to make such an idea work. (iii) We point out that the existence of a quasi-invariant measure is a qualitative statement, showing a delicate persistency property of the dynamics. In particular, this persistence property due to the quasi-invariance is stronger than the (usual) persistence of regularity. In a future work, we plan to construct Hamiltonian dynamics possessing the persistence of regularity such that the Gaussian measure μ s and the transported measure under the dynamics are mutually singular. Remark 1.4 Let us briefly discuss the situation for the related cubic (second order) NLS: i t u = 2 x u ± u 2 u, (x, t) T R. (1.7) It is known to be completely integrable and possesses an infinite sequence of conservation laws H k, k N {0}, controlling the H k -norm [1,2,35]. Associated to the conservation laws H k, k 1, there exists an infinite sequence of invariant weighted Gaussian measures ρ k supported on H k 1 2 ε (T), ε>0[8,74]. As mentioned above, one may combine this invariance and the mutual absolute continuity of ρ k and the Gaussian measure μ k to deduce quasi-invariance of μ k under the dynamics of (1.7), k 1. It may be of interest to investigate quasi-invariance of μ s for non-integer values of s. 1.4 Organization of the paper In Sect. 2, we introduce some notations. In Sect. 3, we apply several transformations to (1.1) and derive a new renormalized equation. We also prove a key factorization lemma (Lemma 3.1) which play a crucial role in the subsequent nonlinear analysis. We then investigate invariance properties of Gaussian measures under several transformations in Sect. 4. In Sect. 5, we prove Theorem 1.2 for s > 1 as a consequence of Ramer s result [62]. By establishing a crucial energy estimate and performing global analysis of truncated measures, we finally present the proof of Theorem 1.2 for the full range s > 3 4 in Sect. 6. In Appendix A, we discuss the well-posedness issue of the Cauchy problem (1.1). Then, we use it to study the approximation property of truncated dynamics in Appendix B, which is used in the proof of Theorem 1.2 in Sect. 6. 2 Notations Given N N, weusep N to denote the Dirichlet projection onto the frequencies { n N} and set P >N := Id P N. Define E N and E N by E N = P N L 2 (T) = span{e inx : n N}, E N = P >N L 2 (T) = span{e inx : n > N}.

Quasi-invariant Gaussian measures for the cubic fourth order 1127 Given s > 1 2,letμ s be the Gaussian measure on L 2 (T) defined in (1.5). Then, we can write μ s as μ s = μ s,n μ s,n, (2.1) where μ s,n and μ s,n are the marginal distributions of μ s restricted onto E N and E N, respectively. In other words, μ s,n and μ s,n are induced probability measures under the following maps: u N : ω u N (x; ω) = g n (ω) n s einx, (2.2) n N u N : ω u N (x; ω) = n >N respectively. Formally, we can write μ s,n and μ s,n as g n (ω) n s einx, (2.3) dμ s,n = Z 1 s,n e 1 2 P N u N 2 H s du N and dμ s,n = Ẑ 1 s,n e 1 2 P >N u N 2 H s du N. (2.4) Given r > 0, we also define a probability measure μ s,r by dμ s,r = Z 1 s,r 1 { v L 2 r}dμ s. (2.5) The defocusing/focusing nature of the Eq. (1.1) does not play any role, and thus we assume that it is defocusing, i.e. with the + sign in (1.1). Moreover, in view of the time reversibility of the equation, we only consider positive times in the following. 3 Reformulation of the cubic fourth order NLS In this section, we apply several transformations to (1.1) and reduce it to a convenient form on which we perform our analysis. Given t R, we define a gauge transformation G t on L 2 (T) by setting G t [ f ]:=e 2it ffl f 2 f, (3.1) where ffl 1 T f (x)dx := 2π T f (x)dx. Given a function u C(R; L2 (T)), we define G by setting G[u](t) := G t [u(t)]. Note that G is invertible and its inverse is given by G 1 [u](t) = G t [u(t)]. Let u C(R; L 2 (T)) beasolutionto(1.1). Define ũ by ũ(t) := G[u](t) = e 2it ffl u(t) 2 u(t). (3.2)

1128 T. Oh, N. Tzvetkov Then, it follows from the the mass conservation that ũ is a solution to the following renormalized fourth order NLS: ) i t ũ = x ( ũ 4 ũ + 2 2 ũ 2 dx ũ. (3.3) T Next, define the interaction representation v of ũ by v(t) = S( t)ũ(t), (3.4) where S(t) = e it 4 x. For simplicity of notations, we use v n to denote the Fourier coefficient of v in the following, when there is no confusion. By writing (3.4) onthe Fourier side, we have v n (t) = e itn4 ũ n (t). (3.5) Then, with (3.5), we can reduce (3.3) to the following equation for {v n } n Z : t v n = ie itn4 (i t ũ n n 4 ũ n ) = i e iφ( n)t v n1 v n2 v n3 + i v n 2 v n Ɣ(n) =: N (v) n + R(v) n, (3.6) where the phase function φ( n) and the plane Ɣ(n) are given by and φ( n) = φ(n 1, n 2, n 3, n) = n 4 1 n4 2 + n4 3 n4 (3.7) Ɣ(n) ={(n 1, n 2, n 3 ) Z 3 : n = n 1 n 2 + n 3 and n 1, n 3 = n}. (3.8) The phase function φ( n) admits the following factorization. Lemma 3.1 Let n = n 1 n 2 + n 3. Then, we have φ( n) = (n 1 n 2 )(n 1 n) ( n 2 1 + n2 2 + n2 3 + n2 + 2(n 1 + n 3 ) 2). (3.9) Proof With n = n 1 n 2 + n 3,wehave { } φ( n) = (n 1 n 2 ) (n 3 1 + n3 2 n3 3 n3 ) + (n 2 1 n 2 + n 1 n 2 2 n2 3 n n 3n 2 ) =: (n 1 n 2 )(I + II). (3.10) On the one hand, we have I = (n 1 n)(n 2 1 + n 1n + n 2 + n 2 2 + n 2n 3 + n 2 3 ). (3.11)

Quasi-invariant Gaussian measures for the cubic fourth order 1129 On the other hand, with n 2 = n 1 + n 3 n, wehave II = 2n 3 1 + 3(n 3 n)n 2 1 + (n2 3 2n 3n + n 2 )n 1 n 2 3 n n 3n 2 = (n 1 n) ( 2n 2 1 + (3n 3 n)n 1 + n 2 3 + n 3n ). (3.12) From (3.10) with (3.10) and (3.12) with n 2 = n 1 + n 3 n, we obtain φ( n) = (n 1 n 2 )(n 1 n)(3n 2 1 + n2 2 + 2n2 3 + n2 + 3n 1 n 3 + n 2 n 3 + n 3 n) = (n 1 n 2 )(n 1 n) ( n 2 1 + n2 2 + n2 3 + n2 + 2(n 1 + n 3 ) 2). In the remaining part of the paper, we present the proof of Theorem 1.2 by performing analysis on (3.6). In view of Lemma 3.1, we refer to the first term N (v) and the second term R(v) on the right-hand side of (3.6) as the non-resonant and resonant terms, respectively. While we do not have any smoothing on R(v) under a time integration, Lemma 3.1 shows that there is a smoothing on the non-resonant term N (v). We will exploit this fact in Sect. 5. In Sect. 6, we will exploit a similar non-resonant behavior in establishing a crucial energy estimate (Proposition 6.1). 4 Gaussian measures under transformations In this section, we discuss invariance properties of Gaussian measures under various transformations. Lemma 4.1 Let t R. Then, the Gaussian measure μ s defined in (1.5) is invariant under the linear map S(t). Proof Note that μ s can be written as an infinite product of Gaussian measures: μ s = n Z ρ n, where ρ n is the probability distribution for û n. In particular, ρ n is a mean-zero Gaussian probability measure on C with variance 2 n 2s. Then, noting that the action of S(t) on û n is a rotation by e itn4, the lemma follows from the rotation invariance of each ρ n. Lemma 4.2 Given a complex-valued mean-zero Gaussian random variable g with variance σ, i.e. g N C (0,σ),letTg = e it g 2 gforsomet R. Then, T g N C (0,σ). Proof By viewing C R 2,letx = (x, y) = (Re g, Im g) and u = (u,v) = (Re Tg, Im Tg). Noting that Tg = g, wehavet 1 g = e it g 2 g. In terms of x and u, wehave x = T 1 u = (u cos t u 2 + v sin t u 2, u sin t u 2 + v cos t u 2 ).

1130 T. Oh, N. Tzvetkov Then, with C t = cos t u 2 and S t = sin t u 2, a direct computation yields det D u T 1 = C t 2tu 2 S t + 2tuvC t S t 2tuvS t + 2tv 2 C t S t 2tu 2 C t 2tuvS t C t 2tuvC t 2tv 2 S t = { Ct 2 2tuvCt 2 2tv 2 S t C t 2tu 2 S t C t + 4t 2 u 3 vs t C t + 4t 2 u 2 v 2 St 2 + 2tuvCt 2 4t 2 u 2 v 2 Ct 2 4t 2 uv 3 } S t C t { St 2 + 2tuvSt 2 2tv 2 S t C t 2tu 2 S t C t + 4t 2 u 3 vs t C t 4t 2 u 2 v 2 Ct 2 2tuvSt 2 + 4t 2 u 2 v 2 St 2 4t 2 uv 3 } S t C t = 1. Let μ and μ be the probability distributions for g and Tg. Then, for a measurable set A C R 2,wehave μ(a) = μ(t 1 A) = 1 πσ = 1 πσ A T 1 A e u 2 σ dudv = μ(a). e x 2 1 σ dxdy = πσ A e T 1 u 2 σ det D u T 1 dudv This proves the lemma. Next, we extend Lemma 4.2 to the higher dimensional setting. Lemma 4.3 Let s R and N N. Then, for any t R, the Gaussian measure μ s,n defined in (2.4) is invariant under the map G t defined in (3.1). While we could adapt the proof of Lemma 4.2 to the higher dimensional setting, this would involve computing determinants of larger and larger matrices. Hence, we present an alternative proof in the following. Proof Given N N,letE N = span{e inx : n N}.Givenu E N,letv(t) = G t [u] for t R. Then, noting that t M(v(t)) = 0, where M(v(t)) = n N v n(t) 2,we see that v n satisfies the following system of ODEs: dv n = 2iM(v)v n dt, n N, (4.1) With a n = Re v n and b n = Im v n, we can rewrite (4.1)as { da n = 2M(v)b n dt n N. (4.2) db n = 2M(v)a n dt, Let L N be the infinitesimal generator for (4.2). Then, μ s,n is invariant under G t for any t R if and only if (L N ) μ s,n = 0. See [44]. Note that the last condition is

Quasi-invariant Gaussian measures for the cubic fourth order 1131 equivalent to (a,b) R 2N+2 L N F(a, b)dμ s,n (a, b) = 0 (4.3) for all test functions F C (R 2N+2 ; R). From(4.2), we have L N F(a, b) = n N ( 2M(a, b)b n + 2M(a, b)a n a n b n ) F(a, b), where M(a, b) = n N (a2 n + b2 n ). Then, by integration by parts, we have (a,b) R 2N+2 L N F(a, b)dμ s,n (a, b) = 2Z 1 N 2Z 1 N = 4Z 1 N 4Z 1 N = 0 n N n N n N n N F(a, b) R2N+2 a n F(a, b) R2N+2 b n ( F(a, b) 1 R2N+2 ( F(a, b) 1 R2N+2 { M(a, b)b n e 2 1 k N { M(a, b)a n e 2 1 k N ) M(a, b) 2 n 2s a n b n e 2 1 k N ) M(a, b) 2 n 2s a n b n e 2 1 k N a 2 k k 2s + b2 k k 2s }dadb a 2 k k 2s + b2 k k 2s }dadb a 2 k k 2s + b2 k k 2s dadb a 2 k k 2s + b2 k k 2s dadb This proves (4.3). In the following, we assume that s > 1 2 such that μ s is a well-defined probability measure on L 2 (T) and G t defined in (3.1) makes sense on supp(μ s ) = L 2 (T). Lemma 4.4 Let s > 1 2. Then, for any t R, the Gaussian measure μ s defined in (1.5) is invariant under the map G t. Note that, when s = 1, Lemma 4.4 basically follows from Theorem 3.1 in [53] which exploits the properties of the Brownian loop under conformal mappings. For general s > 2 1, such approach does not seem to be appropriate. In the following, we present the proof, using Lemma 4.3. Proof Fix t R.GivenN N,letF N C b (L 2 (T); R) be a test function depending only on the frequencies { n N}. Then, we claim that F N G t (u)dμ s (u) = F N (u)dμ s (u). (4.4) L 2 L 2

1132 T. Oh, N. Tzvetkov With a slight abuse of notations, we write F N (u) = F ( {u n } n N ) = FN (u N, u N+1,...,u N 1, u N ). (4.5) Let v = G t [u], where u is as in (1.6). Then, we have v n = e 2it k Z g k 2 k 2s g g n k 2 n s = e2it k >N k 2s e 2it g k 2 k N k 2s g n n s. By the independence of {g n } n N and {g n } n >N, we can write = 0 1 such that g n (ω) = { g n (ω 0 ), ω 0 0, if n N, g n (ω 1 ), ω 1 1, if n > N. Then, we have F N G t (u)dμ s (u) = I N (ω 1 )dp(ω 1 ), (4.6) L 2 1 where I N (ω 1 ) is given by I N (ω 1 ) = 0 F N ({ e 2it g k (ω 1 ) 2 k >N k 2s e 2it g k (ω 0 ) 2 } k N g k 2s n (ω 0 ) n s g k (ω 1 ) 2 k 2s n N ) dp(ω 0 ). (4.7) Since s > 2 1,wehaveμ(ω 1) := k >N < almost surely. For fixed ω 1 1, define { g ω 1 n } n N by setting g ω 1 n = e 2itμ(ω1) g n, n N. Then, by the rotational invariance of the standard complex-valued Gaussian random variables and independence of {g n } n N and {g n } n >N, we see that, for almost every ω 1 1, { g ω 1 n } n N is a sequence of independent standard complex-valued Gaussian random variables (in ω 0 0 ). In particular, the law of { g ω 1 n } n N is the same as that of {g n } n N, almost surely in ω 1 1. Then, from the definitions of μ s,n and G t,we can rewrite (4.7) as ({ I N (ω 1 ) = F N e 2it k N 0 ({ = F N e 2it g k (ω 0 ) 2 k N 0 = F N (G t u N )dμ s,n (u N ) E N g ω } 1 k 2s n (ω 0 ) n s } g k 2s n (ω 0 ) n s g ω 1 k (ω 0 ) 2 n N n N ) dp(ω 0 ) ) dp(ω 0 )

Quasi-invariant Gaussian measures for the cubic fourth order 1133 for almost every ω 1 1, where u N = P N u is as in (2.2). Then, it follows from Lemma 4.3 with (4.5) and (2.1) that I N (ω 1 ) = F N (G t u N )dμ s,n (u N ) = F N (u N )dμ s,n (u N ) = F E N E N L 2 N (u)dμ s (u), (4.8) for almost every ω 1 1. Note that the right-hand side of (4.8) is independent of ω 1 1. Therefore, from (4.6) and (4.8), we have F N G t (u)dμ s (u) = F N (u)dμ s (u)dp(ω 1 ) = F N (u)dμ s (u) L 2 L 2 L 2 1 This proves (4.4). Next, given F C b (L 2 (T); R), letf N (u) = F(P N u), N N. Then, F N (u) converges to F(u) almost surely with respect to μ s.also,f N (G t u) converges to F(G t u) almost surely with respect to μ s. Then, from the dominated convergence theorem and (4.4), we have F G t (u)dμ s (u) = lim F N G t (u)dμ s (u) = lim F N (u)dμ s (u) L 2 N L 2 N L 2 = F(u)dμ s (u) L 2 for all F C b (L 2 (T); R). Hence, the lemma follows (see, for example, [26, Proposition 1.5]). Lastly, we conclude this section by stating the invariance property of quasiinvariance under a composition of two maps. Lemma 4.5 Let (X,μ) be a measure space. Suppose that T 1 and T 2 are maps on X into itself such that μ is quasi-invariant under T j for each j = 1, 2. Then, μ is quasi-invariant under T = T 1 T 2. Proof Suppose that A X is a measurable set such that μ(a) = 0. By the quasiinvariance of μ under T 1, this is equivalent to μ(t1 1 A) = 0. Then, by the quasiinvariance of μ under T 2, the pushforward measure T μ satisfies T μ(a) = μ(t 1 A) = μ ( T 1 2 (T 1 1 A) ) = 0. Conversely, if T μ(a) = 0, then we have μ(t1 1 A) = 0, which in turn implies μ(a) = 0. Hence, μ and T μ are mutually absolutely continuous. 5 Ramer s argument: s > 1 In this section, we present the proof of Theorem 1.2 for s > 1. Our basic approach is to apply Ramer s result after exhibiting a sufficient smoothing on the nonlinear part.

1134 T. Oh, N. Tzvetkov As it is written, the Eq. (1.1) or(3.6) does not manifest a smoothing in an explicit manner. In the following, we perform a normal form reduction and establish a nonlinear smoothing by exploiting the dispersion of the equation. 5.1 Normal form reduction By writing (3.6) in the integral form, we have v n (t) = v n (0) i t 0 e iφ( n)t v n1 v n2 v n3 (t )dt + i Ɣ(n) t 0 v n 2 v n (t )dt =: v n (0) + N(v)(n, t) + R(v)(n, t). (5.1) Lemma 3.1 states that we have a non-trivial (in fact, fast) oscillation caused by the phase function φ( n) in the non-resonant part N(v). The main idea of a normal form reduction is to transform the non-resonant part N(v) into smoother terms of higher degrees, exploiting this rapid oscillation. More concretely, integrating by parts, we formally have N(v)(n, t) = Ɣ(n) e iφ( n)t φ( n) v n1 (t )v n2 (t )v n3 (t t ) t =0 Ɣ(n) t e iφ( n)t 0 φ( n) = φ( n) v n 1 (t)v n2 (t)v n3 (t) 1 φ( n) v n 1 (0)v n2 (0)v n3 (0) Ɣ(n) Ɣ(n) t e iφ( n)t { } 2 N (v)n1 + R(v) n1 vn2 v n3 (t )dt 0 φ( n) Ɣ(n) t e iφ( n)t { } v n1 N (v)n2 + R(v) n2 vn3 (t )dt 0 φ( n) Ɣ(n) e iφ( n)t t (v n1 v n2 v n3 )(t )dt =: I + II + III + IV. (5.2) In view of Lemma 3.1, the phase function φ( n) appearing in the denominators allows us to exhibit a smoothing in N(v). See Lemma 5.1 below. At this point, the computation in (5.2) is rather formal and thus requires justification in several steps. In the first step, we switched the order of the time integration and the summation: t i e iφ( n)t v n1 v n2 v n3 (t )dt = i t e iφ( n)t v n1 v n2 v n3 (t )dt. (5.3) 0 Ɣ(n) Ɣ(n) 0 With w = F 1( v n ) = n Z v n e inx,wehave v n1 v n2 v n3 w 3 L 3 w 3 H 1 6 Ɣ(n) = v 3 H 1. (5.4) 6

Quasi-invariant Gaussian measures for the cubic fourth order 1135 Hence, the sum Ɣ(n) e iφ( n)t v n1 v n2 v n3 (t ) is absolutely convergent with a bound uniform in time t, provided that v C(R; H s (T)) with s 1 6. This justifies (5.3). If v C(R; H s (T)) with s 1 6, it follows from (3.6) and a computation similar to (5.4) that v n C 1 (R). This allows us to apply integration by parts and the product rule. Lastly, we need to justify the switching of the time integration and the summation in the last equality of (5.2). By crudely estimating with (3.6), (5.4) and Lemma 3.1 (note that φ( n) 1onƔ(n)), we have e iφ( n)t { } N (v)n1 + R(v) n1 vn2 v n3 (t ) φ( n) N (v) v n2 v n3 n 1 + R(v) n1 l n1 n Ɣ(n) n 2,n 3 2 n 3 v(t ) 3 H 1 v(t ) 2 L 6 2. (5.5) Hence, the series on the left-hand side of (5.5) is absolutely convergent with a bound uniform in time t, provided that v C(R; H s (T)) with s 1 6. This justifies the last equality in (5.2). The following lemma shows a nonlinear smoothing for (3.6). Note that the amount of smoothing for R(v) depends on the regularity s > 2 1. Lemma 5.1 Let s > 2 1. Then, we have N(v)(t) H s+2 v(0) 3 H s + v(t) 3 H s + t R(v)(t) H 3s t sup v(t ) 5 H s, (5.6) t [0,t] sup v(t ) 3 H s. (5.7) t [0,t] Proof By Lemma 3.1 and the algebra property of H s (T), s > 1 2,wehave I H s+2 n s v n1 (t)v n2 (t)v n3 (t) n s Ɣ(n) l 2 n = {F 1 ( v n )(t)} 3 H s v(t) 3 H s. n=n 1 n 2 +n 3 j=1 3 v n j (t) l 2 n The second term II in (5.2) can be estimated in an analogous manner. Similarly, by Lemma 3.1, (3.6), and the algebra property of H s (T), s > 1 2,wehave III H s+2 t sup t [0,t] = t sup t [0,t] n s { } N (v) n1 + R(v) n1 vn2 v n3 (t ) Ɣ(n) {F 1 ( v n )(t )} 5 H s t l 2 n sup v(t ) 5 t H s. [0,t]

1136 T. Oh, N. Tzvetkov The fourth term IV in (5.2) can be estimated in an analogous manner. From (5.1) and l 2 n l6 n,wehave R(v)(t) H 3s t t ( ) 1 sup n 6s v n (t ) 6 2 = t sup n s v n (t ) 3 t [0,t] n Z t l 6 n [0,t] sup n s v n (t ) 3 t l 2 = t sup v(t ) 3 n [0,t] t H s. [0,t] This proves the second estimate (5.7). 5.2 Consequence of Ramer s result In this subsection, we present the proof of Theorem 1.2 for s > 1. The main ingredient is Ramer s result [62] along with the nonlinear smoothing discussed in the previous subsection. We first recall the precise statement of the main result in [62] for readers convenience. Proposition 5.2 (Ramer [62]) Let (i, H, E) be an abstract Wiener space and μ be the standard Gaussian measure on E. Suppose that T = Id + K : U Ebea continuous (nonlinear) transformation from some open subset U E into E such that (i) T is a homeomorphism of U onto an open subset of E. (ii) We have K (U) H and K : U H is continuous. (iii) For each x U,themap DK(x) is a Hilbert Schmidt operator on H. Moreover, DK: x U DK(x) HS(H) is continuous. (iv) Id H + DK(x) GL(H) for each x U. Then, μ and μ T are mutually absolutely continuous measures on U. Here, HS(H) denotes the space of Hilbert Schmidt operators on H and GL(H) denotes invertible linear operators on H with a bounded inverse. Given t,τ R, let (t): L 2 L 2 be the solution map for (1.1) and (t,τ): L 2 L 2 be the solution map for (3.6), 2 sending initial data at time τ to solutions at time t. When τ = 0, we may denote (t, 0) by (t) for simplicity. By inverting the transformations (3.2) and (3.4) with (5.1), we have (t)(u 0 ) = G 1 S(t) (t)(u 0 ) = G 1 S(t)(u 0 + N(v)(t) + R(v)(t)), (5.8) where v(t) = (t)(u 0 ) and N is given by (5.2). Now, write (t) = Id + K (t), where K (t)(u 0 ) := N(v)(t) + R(v)(t) 2 Note that (3.6) is non-autonomous. We point out that this non-autonomy does not play an essential role in the remaining part of the paper, since all the estimates hold uniformly in t R.

Quasi-invariant Gaussian measures for the cubic fourth order 1137 and v is the solution to (3.6) with v t=0 = u 0. In view of Lemmas 4.1, 4.4, and 4.5,it suffices to show that μ s is quasi-invariant under (t). Fix s > 1 and σ 1 > 1 2 sufficiently close to 1 2. First, note that μ s is a probability measure on H s σ 1(T). GivenR > 0, let B R be the open ball of radius R centered at the origin in H s σ 1(T). The following proposition shows that the hypotheses of Ramer s result in [62] are indeed satisfied. Proposition 5.3 Let s > 1. Given R > 0, there exists τ = τ(r) >0 such that, for each t (0,τ(R)], the following statements hold: (i) (t) is a homeomorphism of B R onto an open subset of H s σ 1(T). (ii) We have K (t)(b R ) H s (T) and K (t) : B R H s (T) is continuous. (iii) For each u 0 B R,themapDK(t) u0 is a Hilbert Schmidt operator on H s (T). Moreover, DK (t) : u 0 B R DK(t) u0 HS(H s (T)) is continuous. (iv) Id H s + DK(t) u0 GL(H s (T)) for each u 0 B R. We first present the proof of Theorem 1.2 for s > 1, assuming Proposition 5.3. Thanks to Ramer s result (Proposition 5.2 above), Proposition 5.3 implies that μ s and the pullback measure (t) μ s := μ s (t) are mutually absolutely continuous as measures restricted to the ball B R for any t (0,τ(R)]. Proof of Theorem 1.2 for s > 1GivenR > 0, let B R H s σ 1 be the open ball of radius R centered at the origin as above. Fix T > 0. It follows from the growth estimate (A.12) oftheh s σ 1-norm that sup v(t) H s σ 1 C(T, R) =: R (5.9) t [0,T ] for all solutions v to (3.6) with v t=0 B R. Suppose that A B H s σ 1 is a Borel set in H s σ 1 such that μ s (A) = 0. Given R > 0, let R be as in (5.9). Then, from (5.9), we have (t)(a B R ) B R (5.10) for all t [0, T ]. Noting that μ s (A B R ) = 0, it follows from Proposition 5.3 and the result in [62] that μ s ( (t)(a B R )) = 0 (5.11) for any 0 t τ, where τ = τ(r ) is as in Proposition 5.3. Inviewof(5.10), we can iteratively apply Proposition 5.3 and the main result in [62] on time interval [ jτ,(j + 1)τ] and see that (5.11) holds for all t [0, T ]. In particular, we have μ s ( (T )(A B R )) = 0. Now, letting R, it follows from the continuity from below of a measure that μ s ( (T )(A)) = 0. Note that the choice of T was arbitrary. In view of the time

1138 T. Oh, N. Tzvetkov reversibility of the Eq. (3.6), we conclude that μ s is quasi-invariant under the flow of (3.6). Therefore, Theorem 1.2 follows from (5.8) with Lemmas 4.1, 4.4, and 4.5. The remaining part of this section is devoted to the proof of Proposition 5.3. The claim (i) follows from the well-posedness of (3.6) inh s σ 1. In particular, the continuity of (t) on H s σ 1 with the time reversibility implies (i). As before, from (A.12), we have the uniform growth bound: sup v(t) H s σ 1 C Rθ τ R =: R τ (5.12) t [0,τ] for all solutions v to (3.6) with v t=0 = u 0 B R. Then, the claim (ii) follows from Lemma 5.1 and the continuity of (t) on H s σ 1. We postpone the proof of the claim (iii) and first prove the claim (iv). For fixed u 0 B R H s σ 1 and t R, define a map F : H s H s by F(h) = (t)(u 0 + h) u 0 = h + K (t)(u 0 + h), h H s. Then, by computing a derivative of F at the origin, we have 3 DF 0 = Id H s +DK(t) u0. This is clearly a linear map. Moreover, the boundedness of DF 0 on H s follows from the claim (iii). Note that F is invertible with the inverse F 1 given by F 1 (h) = ( t)(u 0 + h) u 0 = h + K ( t)(u 0 + h). Hence, it follows from the chain rule that DF 0 = Id H s + DK(t) u0 is invertible. Moreover, we have (DF 0 ) 1 = Id H s + DK( t) u0. (5.13) Hence, we proved the claim (iv) except for the boundedness of (DF 0 ) 1. We will prove the boundedness of (DF 0 ) 1 at the end of this section. Next, we prove the claim (iii). In the following, we will prove that DK(t) u0 is Hilbert-Schmidt on H s for u 0 B R H s σ 1 as long as t = t(r) 1. Given u 0 B R H s σ 1,letv be the global solution to (3.6) with v t=0 = u 0. We first introduce some notations. Given a multilinear 4 expression M(v,v,...,v), we use M(v,v,...,v ) to denote the sum of the form M(w,v,...) + M (v,w,v,...), where each multilinear term has exactly one factor of w and the remaining arguments are v. For example, we have v n 1 (t)v n 2 (t)v n 3 (t) = w n1 (t)v n2 (t)v n3 (t) + v n1 (t)w n2 (t)v n3 (t) + v n1 (t)v n2 (t)w n3 (t). We use a similar convention for multilinear expressions in v(0). In this case, we use M(v (0), v (0),...) to denote the sum of the form M(w(0), v(0),...) + 3 By viewing (H s, H s σ 1,μ s ) as an abstract Wiener space, DF 0 is the so-called H-derivative of F at 0, where H = H s is the Cameron Martin space. See [62]. 4 By multilinearity, we mean it is either linear or conjugate linear in each argument, i.e. linear over real numbers.

Quasi-invariant Gaussian measures for the cubic fourth order 1139 M(v(0), w(0), v(0),...), where each multilinear term has exactly one factor of w(0) and the remaining arguments are v(0). Let w(t) be a solution to the following linear equation: t w n = i e iφ( n)t vn 1 vn 2 vn 3 + i vn 2 vn Ɣ(n) (5.14) w t=0 = w(0). Given (m 1, m 2, m 3 ) Z 3 and n Z, we use the following shorthand notation: ( m, n) := (m 1, m 2, m 3, n). (5.15) Then, by a direct computation with (3.6), (5.1), and (5.2), we have F [ DK(t) u0 (w(0)) ] (n) = i + Ɣ(n) + 2i 2i i + i t 0 t t 0 t 0 0 e iφ( n)t t 0 v n (t ) 2 v n (t )dt φ( n) v n 1 (t)vn 2 (t)vn 3 (t) 1 φ( n) v n 1 (0)vn 2 (0)vn 3 (0) Ɣ(n) e iφ( n)t iφ( m,n 1 )t vm 1 vm 2 vm 3 vn 2 vn 3 (t )dt (n 1,n 2,n 3 ) Ɣ(n) (m 1,m 2,m 3 ) Ɣ(n 1 ) Ɣ(n) e iφ( n)t φ( n) (n 1,n 2,n 3 ) Ɣ(n) (m 1,m 2,m 3 ) Ɣ(n 2 ) Ɣ(n) e iφ( n)t φ( n) φ( n) v n 1 2 v n 1 v n 2 v n 3 (t )dt e iφ( n)t +iφ( m,n 2 )t φ( n) v n 1 v n 2 2 v n 2 v n 3 (t )dt, where φ( n) and Ɣ(n) are as in (3.7) and (3.8), respectively. Fix σ 2 > 2 1 (to be chosen later) and write DK(t) u0 (w(0)) = x σ 2 A t (w(0)) where A t (w(0)) is given by F [ A t (w(0)) ] (n) = i + Ɣ(n) +2i t 0 e iφ( n)t t 0 n σ 2 v n (t ) 2 v n (t )dt v n 1 v m 1 v m 2 v m 3 v n 3 (t )dt φ( n) n σ 2 vn 1 (t)vn 2 (t)vn 3 (t) 1 φ( n) n σ 2 vn 1 (0)vn 2 (0)vn 3 (0) Ɣ(n) e iφ( n)t iφ( m,n 1 )t n σ 2 vm φ( n) 1 vm 2 vm 3 vn 2 vn 3 (t )dt (n 1,n 2,n 3 ) Ɣ(n) (m 1,m 2,m 3 ) Ɣ(n 1 )

1140 T. Oh, N. Tzvetkov 2i i +i t t 0 t 0 0 Ɣ(n) e iφ( n)t φ( n) (n 1,n 2,n 3 ) Ɣ(n) (m 1,m 2,m 3 ) Ɣ(n 2 ) Ɣ(n) e iφ( n)t φ( n) n σ 2 v n 1 2 v n 1 v n 2 v n 3 (t )dt e iφ( n)t +iφ( m,n 2 )t φ( n) n σ 2 v n 1 v m 1 v m 2 v m 3 v n 3 (t )dt n σ 2 v n 1 v n 2 2 v n 2 v n 3 (t )dt, (5.16) Note that x σ 2 is a Hilbert Schmidt operator on H s. Thus, if we prove that A t is bounded on H s, then it follows that DK(t) u0 is Hilbert-Schmidt on H s. Hence, we focus on proving the boundedness of A t on H s in the following. Let t [0, 1]. Givens > 1, choose σ 1,σ 2 > 2 1 such that s σ 1 > 1 2, s+σ 2 3 s σ 1, and s + σ 2 2 s σ 1. (5.17) Applying Young s inequality and l 2 n l6 n to (5.16) with Lemma 3.1 and (5.17), we have A t (w(0)) H s { sup v(t ) 2 t [0,t] H s+σ 2 w(t ) s+σ 2 3 H 3 + v(t ) 2 H s σ 1 w(t ) H s σ 1 + v(t ) 4 H s σ 1 w(t ) H s σ 1 }. (5.18) Given τ>0, it follows from (A.12) that Then, from (5.14) and (5.19) with (5.17), we have sup v(t) H s σ 1 C(R). (5.19) t [0,τ] sup w(t) H s σ 1 w(0) H s + Cτ sup v(t) 2 H s σ w(t) 1 H s σ 1 t [0,τ] t [0,τ] w(0) H s + C(R)τ sup w(t) H s σ 1. t [0,τ] In particular, by choosing τ = τ(r) >0 sufficiently small, we obtain sup w(t ) H s σ 1 w(0) H s. (5.20) t [0,τ] Finally, it follows from (5.18), (5.19), and (5.20) with (5.17) that A t (w(0)) H s C(R) w(0) H s.

Quasi-invariant Gaussian measures for the cubic fourth order 1141 Therefore, A t is bounded on H s and hence DK(t) u0 is a Hilbert Schmidt operator on H s for all t [0,τ]. The second claim in (iii) basically follows from the continuous dependence of (3.6) and (5.14) (inv) and thus we omit details. It remains to prove the boundedness of (DF 0 ) 1 = (Id H s + DK(t) u0 ) 1 By the time reversibility of the equation and (5.13), the argument above shows that (DF 0 ) 1 Id H s is Hilbert Schmidt on H s by choosing τ = τ(r) sufficiently small. In particular, (DF 0 ) 1 is bounded on H s. This completes the proof of Proposition 5.3. Remark 5.4 The condition s > 1 is necessary for this argument. In estimating the resonant term, i.e. the first term in (5.16) bytheh s σ 1-norms of its arguments, we need to use the second condition s+σ 2 3 s σ 1 in (5.17). Thus, we must have s 3σ 1 + σ 2 2 > 1, since σ 1,σ 2 > 1 2. 6 Proof of Theorem 1.2: s > 3 4 In this section, we present the proof of Theorem 1.2 for s > 3 4. The basic structure of our argument follows the argument introduced in [70] by the second author in the context of the (generalized) BBM equation, with one importance difference. While the energy estimate in [70] was carried out on the H s -norm of solutions (to the truncated equations), we carry out our energy estimate on a modified energy. This introduction of a modified energy is necessary to exhibit a hidden nonlinear smoothing, exploiting the dispersion of the equation. See Proposition 6.1 below. This, in turn, forces us to work with the weighted Gaussian measure ρ s,n,r,t and ρ s,r,t adapted to this modified energy, instead of the Gaussian measure μ s,r with an L 2 -cutoff. See (6.21) and (6.22) below for the definitions of ρ s,n,r,t and ρ s,r,t. Lastly, we point out that this usage of the modified energy is close to the spirit of higher order modified energies in the I -method introduced by Colliander Keel Staffilani Takaoka Tao [22,23]. As in Sect. 5, we carry out our analysis on (3.6). Let us first introduce the following truncated approximation to (3.6): t v n = 1 n N { i where Ɣ N (n) is defined by Ɣ N (n) e iφ( n)t v n1 v n2 v n3 + i v n 2 v n }, (6.1) Ɣ N (n) = Ɣ(n) {(n 1, n 2, n 3 ) Z 3 : n j N} ={(n 1, n 2, n 3 ) Z 3 : n = n 1 n 2 + n 3, n 1, n 3, = n, and n j N}. (6.2)

1142 T. Oh, N. Tzvetkov A major part of this section is devoted to the study of the dynamical properties of (6.1). Note that (6.1) is an infinite dimensional system ODEs for the Fourier coefficients {v n } n Z, where the flow is constant on the high frequencies { n > N}. We also consider the following finite dimensional system of ODEs: t v n = i e iφ( n)t v n1 v n2 v n3 + i v n 2 v n, n N. (6.3) Ɣ N (n) Given t,τ R, denote by N (t,τ)and N (t,τ)the solution maps of (6.1) and (6.3), sending initial data at time τ to solutions at time t, respectively. For simplicity, we set N (t) = N (t, 0) and N (t) = N (t, 0) (6.4) when τ = 0. Then, we have the following relations: N (t,τ)= N (t,τ)p N + P >N and P N N (t,τ)= N (t,τ)p N. (6.5) 6.1 Energy estimate In this subsection, we establish a key energy estimate. Before stating the main proposition, let us first perform a preliminary computation. Given a smooth solution u to (1.1), let v be as in (3.4). Then, from (3.6), we have d dt u(t) 2 H s = d dt v(t) 2 H s = 2Re i e iφ( n)t n 2s v n1 v n2 v n3 v n. (6.6) n Z Ɣ(n) Then, differentiating by parts, i.e. integrating by parts without an integral sign, 5 we obtain d dt v(t) 2 H s = 2Re d [ e iφ( n)t ] dt φ( n) n 2s v n1 v n2 v n3 v n 2Re n Z n Z Ɣ(n) Ɣ(n) e iφ( n)t φ( n) n 2s t (v n1 v n2 v n3 v n ). (6.7) This motivates us to define the following quantity. Given s > 2 1, define the modified energy E t (v) by E t (v) = v 2 H s 2Re n Z Ɣ(n) e iφ( n)t φ( n) n 2s v n1 v n2 v n3 v n =: v 2 H s + R t(v). (6.8) Then, we have the following energy estimate. 5 This is indeed a normal form reduction applied to the evolution Eq. (6.6) for v(t) 2 H s. Compare this with the normal form reduction argument in Sect. 5.

Quasi-invariant Gaussian measures for the cubic fourth order 1143 Proposition 6.1 Let s > 4 3. Then, for any sufficiently small ε>0, there exist small θ>0and C > 0 such that d dt E t(p N v) C v 4+θ v 2 θ L 2 H s 2 1 ε, (6.9) for all N N and any solution v to (6.1), uniformly in t R. Recall that the probability measures μ s and μ s,r defined in (1.5) and (2.5) are supported on H s 1 2 ε (T) for any ε>0, while we have v L 2 r in the support of μ s,r. Before proceeding to the proof of this proposition, recall the following arithmetic fact [39]. Given n N, the number d(n) of the divisors of n satisfies for any δ>0. d(n) C δ n δ (6.10) Proof Let v be a solution to (6.1). Then, from (6.7) and (6.8) with (6.1), we have d dt E t(p N v) = N 1 (v) + R 1 (v) + N 2 (v) + R 2 (v) + N 3 (v) + R 3 (v), (6.11) where N j (v) and R j (v), j = 1, 2, 3, are defined by e iφ( n)t iφ( m,n 1)t N 1 (v)(t) := 4Re i n 2s v m1 v m2 v m3 v n2 v n3 v n φ( n) n N (n 1,n 2,n 3 ) Ɣ N (n) (m 1,m 2,m 3 ) Ɣ N (n 1 ) e iφ( n)t R 1 (v)(t) := 4Re i n 2s v n1 2 v n1 v n2 v n3 v n φ( n) n N Ɣ N (n) e iφ( n)t+iφ( m,n 2)t N 2 (v)(t) := 2Re i n 2s v n1 v m1 v m2 v m3 v n3 v n φ( n) n N (n 1,n 2,n 3 ) Ɣ N (n) (m 1,m 2,m 3 ) Ɣ N (n 2 ) e iφ( n)t R 2 (v)(t) := 2Re i n 2s v n1 v n2 2 v n2 v n3 v n φ( n) n N Ɣ N (n) e iφ( n)t+iφ( m,n)t N 3 (v)(t) := 2Re i n 2s v n1 v n2 v n3 v m1 v m2 v m3 φ( n) n N (n 1,n 2,n 3 ) Ɣ N (n) (m 1,m 2,m 3 ) Ɣ N (n) e iφ( n)t R 3 (v)(t) := 2Re i n 2s v n1 v n2 v n3 v n 2 v n. (6.12) φ( n) n N Ɣ N (n) Here, ( m, n) = (m 1, m 2, m 3, n) and ( m, n j ) = (m 1, m 2, m 3, n j ) are as in (5.15). For simplicity of the presentation, we drop the restriction on the summations in (6.12) with the understanding that v n = 0for n > N. Moreover, we can assume that all

1144 T. Oh, N. Tzvetkov the Fourier coefficients are non-negative. In the following, we establish uniform (in t) estimates for these multilinear terms N j and R j, j = 1, 2, 3. For simplicity, we suppress the t-dependence with the understanding that all the estimates hold with implicit constants independent of t R. Given n,μ Z, define Ɣ(n,μ)by Ɣ(n,μ):= Ɣ(n) {(n 1, n 2, n 3 ) Z 3 : μ = (n n 1 )(n n 3 )} ={(n 1, n 2, n 3 ) Z 3 : n = n 1 n 2 + n 3, n 1, n 3 = n, μ= (n n 1 )(n n 3 )}. Then, given δ>0, it follows from the divisor counting estimate (6.10) that #Ɣ(n,μ)= 1 C δ μ δ. (6.13) Ɣ(n,μ) In the following, we use (6.13) to estimate N j (v) and R j (v), j = 1, 2, 3. For simplicity of the presentation, we drop multiplicative constants depending on δ>0. We now estimate N 1 (v). We first consider the case s < 1. By Sobolev s inequality and interpolation, we have v m1 v m2 v m3 l F 1 ( v n ) 3 L 3 v 3 H 1 v 1+θ v 2 θ (6.14) 6 L 2 Ɣ(n 1 ) n1 H 1 4 +γ for small γ>0and some θ = θ(γ) > 0. Then, by Lemma 3.1 and Cauchy Schwarz inequality (in n and then in n 2, n 3 ) with (6.14), we have N 1 (v) 1 μ n 2 2s v n2 v n3 v n v m1 v m2 v m3 l n Z μ =0 Ɣ(n,μ) max Ɣ(n 1 ) n1 v 2+θ v 2 θ L 2 H 1 4 +γ { ( 1 μ 1+2δ n Z μ =0 Ɣ(n,μ) )( vn 2 1 2 vn 2 )} 2 1 3 (n 2 n 3 )(n n 3 ) 1 2δ n 4 4s max Ɣ(n) (6.15) for small γ,δ > 0 such that 5 4s 2δ >1, where n max := max( n, n 1, n 2, n 3 ). From the divisor counting argument (6.13), we have { ( N 1 (v) v 2+θ v 2 θ 1 L 2 H 4 1 +γ μ n Z μ =0 1+2δ μ δ )( Ɣ(n) { v 2+θ v 2 θ v L 2 H 4 1 +γ n 2 2 v 2 1 n 3 n n n 2,n 3 Z n =n 3 1 2δ n 4 4s 3 vn 2 2 vn 2 )} 2 1 3 (n n 1 )(n n 3 ) 1 2δ n 4 4s max v 4+θ v 2 θ (6.16) L 2 H s 2 1 ε for sufficiently small δ, ε, γ > 0, provided that s > 3 4. } 1 2

Quasi-invariant Gaussian measures for the cubic fourth order 1145 Next, we consider the case s 1. Note that for (n 1, n 2, n 3 ) Ɣ(n), wehave max j=1,2,3 n j n. First, suppose that n 1 n. In this case, we use the fact that max j=1,2,3 m j n 1 for (m 1, m 2, m 3 ) Ɣ(n 1 ). Without loss of generality, assume that m 1 n 1 ( n ). Proceeding as in (6.15) and (6.16) with (6.14), Cauchy-Schwarz inequality, (6.13), and interpolation, we have N 1 (v) n Z μ =0 Ɣ(n,μ) v H s 1+δ v H s 6 5 +δ v 2 H 1 6 { ( 1 μ 1+2δ n Z μ =0 1 μ n 2δ v n 2 v n3 n s 1+δ v n 1 Ɣ(n 1 ) m s 1+δ v m1 v m2 v m3 l n1 Ɣ(n,μ) v H s 1+δ v H s 6 5 +δ v 2 H 1 6 { ( n Z μ =0 )( 1 Ɣ(n) )( 1 μ δ μ 1+2δ Ɣ(n) { v H s 1+δ v H s 5 6 +δ v 2 H 1 6 n 2,n 3 Z v 2 L 2 v H s 1+δ v H s 5 6 +δ v 2 H 1 6 vn 2 2 vn 2 )} 2 1 3 (n n 1 )(n n 3 ) 1 2δ n 4δ vn 2 2 vn 2 )} 2 1 3 (n n 1 )(n n 3 ) 1 2δ n 4δ vn 2 2 vn 2 1 3 n n n =n 3 1 2δ n 4δ 3 } 1 2 v 4+θ v 2 θ (6.17) L 2 H s 2 1 ε for sufficiently small δ, ε > 0 and some θ = θ(s,δ,ε)>0. Suppose that n 1 n. In this case, we have max( n 2, n 3 ) n. Without loss of generality, assume that n 2 n. Proceeding as in (6.15) and (6.16) with (6.13) and (6.14), we have N 1 (v) n Z μ =0 Ɣ(n,μ) v H s 1+δ v 3 H 1 6 v H s 1+δ v 3 H 1 6 1 μ n 2δ n 2 s 1+δ v n2 v n3 n s 1+δ v n v m1 v m2 v m3 l Ɣ(n 1 ) n1 { ( )( 1 n 2 2(s 1+δ) vn 2 2 vn 2 3 μ δ μ 1+2δ (n n μ =0 1 )(n n 3 ) 1 2δ n 4δ Ɣ(n) n Z { n 2,n 3 Z v L 2 v 2 H s 1+δ v 3 H 1 6 n 2 2(s 1+δ) v 2 n 2 v 2 n 3 n n 3 1 2δ n 4δ n =n 3 1 } 1 2 )} 1 2 v 4+θ v 2 θ (6.18) L 2 H s 2 1 ε for sufficiently small δ, ε > 0 and some θ = θ(s,δ,ε)>0. Noting that μ = (n n 1 )(n n 3 ) = (n 2 n 1 )(n 2 n 3 ) under n = n 1 n 2 + n 3, we can estimate N 2 (v) and N 3 (v) in a similar manner.

1146 T. Oh, N. Tzvetkov Next, we estimate R 1 (v). The remaining terms R 2 (v) and R 3 (v) can be estimated in a similar manner. Without loss of generality, suppose that n 3 n. From Lemma 3.1 and the divisor counting argument (6.13), we have { ( )( 1 vn 6 R 1 (v) v H s 1+δ μ δ 1 vn 2 2 n 3 2(s 1+δ) v 2 )} 1 n 2 3 μ1+2δ (n n n Z μ =0 Ɣ(n) 1 )(n n 3 ) 1 2δ n 4δ v 4 L 2 v 2 H s 1+δ v 4+θ L 2 v 2 θ (6.19) H s 2 1 ε for sufficiently small δ, ε > 0 and some θ = θ(s,δ,ε)>0. Therefore, (6.9) follows from (6.16), (6.17), (6.18), and (6.19). This completes the proof of Proposition 6.1. 6.2 Weighted Gaussian measures Our main goal in this subsection is to define weighted Gaussian measures adapted to the modified energy E t (P N v) and E t (v) defined in the previous section. Given N N and r > 0, define F N,r,t (v) and F r,t (v) by F N,r,t (v) = 1 { v L 2 r}e 1 2 R t (P N v) and F r,t (v) = 1 { v L 2 r}e 1 2 R t (v), (6.20) where R t is defined in (6.8). Then, we would like to construct probability measures ρ s,n,r,t and ρ s,r,t of the form: 6 and dρ s,n,r,t = Zs,N,r 1 1 { v L 2 r}e 1 2 E t (P N v) dv = Zs,N,r 1 F N,r,tdμ s (6.21) dρ s,r,t = Zs,r 1 1 { v L 2 r}e 1 2 E t (v) dv = Zs,r 1 F r,tdμ s. (6.22) The following proposition shows that they are indeed well defined probability measures on H s 1 2 ε (T), ε>0. Proposition 6.2 Let s > 1 2 and r > 0. Then, F N,r,t(v) L p (μ s ) for any p 1 with a uniform bound in N N and t R, depending only on p 1 and r > 0. Moreover, for any finite p 1, F N,r,t (v) converges to F r,t (v) in L p (μ s ), uniformly in t R,as N. 6 The normalizing constants Z s,n,r and Z s,r aprioridependont R. It is, however, easy to see that they are indeed independent of t R by (i) noticing that R t (v) defined in (6.8) is autonomous in terms of ũ(t) = S(t)v(t) and (ii) the invariance of μ s under S(t) (Lemma 4.1).