Asymptotic Dynamics of Nonlinear Schrödinger Equations: Resonance-Dominated and Dispersion-Dominated Solutions TAI-PENG TSAI AND HORNG-TZER YAU Courant Institute Abstract We consider a linear Schrödinger equation with a nonlinear perturbation in R3 . Assume that the linear Hamiltonian has exactly two bound states and its eigenvalues satisfy some resonance condition. We prove that if the initial data is sufficiently small and is near a nonlinear ground state, then the solution approaches to certain nonlinear ground state as the time tends to infinity. Furthermore, the difference between the wave function solving the nonlinear Schrödinger equation and its asymptotic profile can have two different types of decay: The resonancedominated solutions decay as t −1/2 or the dispersion-dominated solutions decay c 2002 John Wiley & Sons, Inc. at least like t −3/2 . Contents 1. Introduction 154 2. Preliminaries 162 3. Main Oscillation Terms 171 4. Outline of Proofs and Basic Estimates 175 5. Estimates of the Dispersive Wave 179 6. Excited-State Equation and Normal Form 182 7. Change of the Mass of the Ground State 191 8. Existence of Dispersion-Dominated Solutions 201 9. Proof of Linear Estimates 205 Bibliography 215 Communications on Pure and Applied Mathematics, Vol. LV, 0153–0216 (2002) c 2002 John Wiley & Sons, Inc. 154 T.-P. TSAI AND H.-T. YAU 1 Introduction Consider the nonlinear Schrödinger equation (1.1) i∂t ψ = (−1 + V )ψ + λ|ψ|2 ψ , ψ(t = 0) = ψ0 , where V is a smooth localized potential, λ is an order-1 parameter, and ψ = ψ(t, x) : R × R3 −→ C is a wave function. Let e0 < 0 be the ground state energy to −1 + V , and denote H1 = −1 + V − e0 . The nonlinear bound states to the Schrödinger equation (1.1) are solutions to the equation (1.2) (−1 + V )Q + λ|Q|2 Q = E Q . They are critical points to the energy functional Z 1 1 1 H[φ] = |∇φ|2 + V |φ|2 + λ|φ|4 d x 2 2 4 2 subject to the constraint of fixed L norm. For each bound state Q = Q E , ψ(t) = Qe−i Et , is a solution to the nonlinear Schrödinger equation. We may obtain a family of such bound states by standard bifurcation theory: For each E sufficiently close to e0 so that E − e0 and λ share the same sign, there is a unique small positive solution Q = Q E to equation (1.2) that decays exponentially as x → ∞; see Lemma 2.1. We call this family the nonlinear ground states and shall refer to it as {Q E } E . Let (1.3) HE = −1 + V − E + λQ 2E . We have HE Q E = 0. Since Q E is small and E is close to e0 , the spectral properties of HE are similar to those of H1 . Suppose the initial data of the nonlinear Schrödinger equation ψ0 is near some Q E . Under rather general conditions, the family of nonlinear ground states is stable in the sense that if inf ψ(t) − Q E ei2 2 2,E L is small for t = 0, it remains so for all t; see, for example, [13] for the case λ < 0. See also [18, 19]. Let k·k L 2 denote a local L 2 norm; a precise choice will be made loc later on. One expects that this difference actually approaches zero when measured by a local L 2 norm, i.e., (1.4) lim inf ψ(t) − Q E ei2 2 = 0 . t→∞ 2,E L loc If −1 + V has only one bound state, it is proven in [15] that the evolution will eventually settle down to some ground state Q E∞ with E ∞ close to E. (See also [9] for another proof using techniques from dynamical systems.) Suppose now that −1 + V has multiple bound states, say, two bound states: a ground state φ0 with eigenvalue e0 and an excited state φ1 with eigenvalue e1 , i.e., H1 φ1 = e01 φ1 where e01 = e1 −e0 > 0. The question is whether the evolution with initial data ψ0 near some Q E will eventually settle down to some ground state Q E∞ DYNAMICS OF NLS 155 with E ∞ close to E. Furthermore, can we characterize the asymptotic evolution? In this paper, we shall answer this question positively in the case of two bound states and estimate precisely the rate of relaxation for a certain class of initial data. We now state the main assumptions of this paper. A SSUMPTION A0: −1 + V acting on L 2 (R3 ) has two simple eigenvalues e0 < e1 < 0 with normalized eigenvectors φ0 and φ1 . A SSUMPTION A1: Resonance condition. Let e01 = e1 − e0 be the spectral gap of the ground state. We assume that 2e01 > |e0 | so that 2e01 is in the continuum spectrum of H1 . Furthermore, for some constant γ0 > 0 and all real s sufficiently small, 1 2 H1 2 (1.5) lim φ0 φ1 , Im P φ0 φ1 ≥ γ0 > 0 . σ →0+ H1 − σ i − 2e01 − s c We shall use 0i to replace σ i and the limit limσ →0+ later on. A SSUMPTION A2: For λQ 2E sufficiently small, the bottom of the continuous spectrum to −1 + V + λQ 2E , 0, is not a generalized eigenvalue; i.e., it is not a resonance. Also, we assume that V satisfies the assumption in Yajima [20] so that the W k, p estimates k ≤ 2 for the wave operator W H = limt→∞ eit H eit (1+E) hold for k ≤ 2; i.e., there is a small σ > 0 such that |∇ α V (x)| ≤ Chxi−5−σ for |α| ≤ 2 . Also, the functions (x · ∇)k V for k = 0, 1, 2, 3, are −1 bounded with a −1 bound less than 1: (x · ∇)k V φ ≤ σ0 k−1φk2 + C kφk2 , σ0 < 1 , k = 0, 1, 2, 3. 2 Assumption A2 contains some standard conditions to assure that most tools in linear Schrödinger operators apply. These conditions are certainly not optimal. The main assumption above is the condition 2e01 > |e0 | in assumption A1. The rest of assumption A1 just consists of generic assumptions. This condition states that the excited state energy is closer to the continuum spectrum than to the ground state energy. It guarantees that twice the excited state energy of H1 (which one obtains from taking the square of the excited state component) becomes a resonance in the continuum spectrum (of H1 ). This resonance produces the main relaxation mechanism. If this condition fails, the resonance occurs in higher-order terms and a proof of relaxation will be much more complicated. Also, the rate of decay will be different. Define the notation p ∼ min εn, n −1 t −1/2 , (1.6) hxi = 1 + x 2 , {t}ε = ε−2 n −2 + 20t , {t}−1/2 ε where n = kψ0 k L 2 and 0 = O(n 2 ) is a positive constant to be specified later in (6.3) of Lemma 6.1. For the moment, we remark that 0 is of order 2λ2 n 2 times the quantity in (1.5). The subscript ε is a small parameter to be specified in Theorems 1.3 and 1.4. We shall often drop it in the proofs of Theorems 1.3 and 1.4. We 156 T.-P. TSAI AND H.-T. YAU denote by L r2 the weighted L 2 spaces (r may be positive or negative), (1.7) L r2 (R3 ) ≡ φ ∈ L 2 (R3 ) : hxir φ ∈ L 2 (R3 ) . Our space for initial data is Y ≡ H 1 (R3 ) ∩ L r20 (R3 ) , (1.8) r0 > 3 . to denote The parameter r0 > 3 is fixed, and we can choose, We shall use say, r0 = 4 for the rest of this paper. Our first theorem states that if the initial data ψ0 is small in Y and the distance between ψ0 and a nonlinear ground state Q ∗ is small, then the solution ψ(t) has to settle down to some asymptotic nonlinear ground state as t → ∞. Furthermore, the difference between ψ0 and the asymptotic nonlinear ground state at t = ∞ is bounded above by the order O(t −1/2 ). Recall that λ = O(1) is fixed. L 2loc L 2−r0 . T HEOREM 1.1 Assume that assumptions A0, A1, and A2 on V hold. Then there are small universal positive constants ε0 and n 0 > 0 such that, for any nonlinear ground state Q ∗ with mass n = kQ ∗ k L 2 < n 0 and any initial data ψ0 satisfying kψ0 − ei20 Q ∗ kY ≤ ε02 n 2 for some 20 ∈ R, there exist an energy E ∞ and a function 2(t) such that kQ E∞ k L 2 − n = O(ε02 n), 2(t) = −E ∞ t + O(log t), and ψ(t) − Q E ei2(t) 2 ≤ C(1 + t)−1/2 . (1.9) ∞ L loc To describe more detailed behavior of the solution ψ(t), we need various spectral properties of the linearized operator. All statements we make here will be proven in Section 2. Let L be the operator obtained from linearizing the Schrödinger equation (1.1) around the trivial solution Qe−i Et (see Section 2), i.e., (1.10) Lh = i −1 (−1 + V − E + λQ 2 )h + λQ 2 (h + h) . Notice that L is not self-adjoint due to the conjugation. If we can decompose h into its real and imaginary parts, the operator L can be written in the matrix forms 0 L− L ←→ −L + 0 where (1.11) L − := −1 + V − E + λQ 2E = HE , L + = L − + 2λQ 2E . Explicitly, we have L( f + ig) = L − g − i L + f . Notice that L is a small perturbation of the Hamiltonian H1 . Hence the spectral properties of L are closely related to those of H1 . Since H1 has two eigenvalues, 0 and e01 , we expect L to have two eigenvalues as well. From (1.2) we have L − Q E = 0. If we differentiate (1.2) with respect to E, we have L + R E = Q E , where R E = ∂ E Q E . Let S E be the space spanned by i Q E and its tangent R E , i.e., (1.12) S = S E = spanR {R E , i Q E } . DYNAMICS OF NLS 157 Clearly, S is the generalized eigenspace of L with eigenvalue 0. L has a pair of eigenvalues ±iκ, where κ is a perturbation of e01 = e1 − e0 of H1 . The corresponding generalized eigenfunctions, u and v, are perturbations of the linear excited state φ1 . They are real-valued functions characterized by (1.13) L + u = κv , L − v = κu , (u, v) = 1 . In particular, for all real numbers α and β, we have L2 (αu + βiv) = −κ 2 (αu + βiv). Thus (1.14) Eκ (L) = spanR {u, iv} is the generalized eigenspace with eigenvalues ±iκ. Finally, the space of the continuous spectrum, Hc (L), is characterized by the relation (1.15) Hc (L) = f + ig ∈ L 2 : f, g real, f ⊥ Q E , v; g ⊥ R E , u . We have the following spectral decomposition result, to be proven in Section 2.2. L EMMA 1.2 (Spectral Decomposition) The space of complex-valued L 2 functions in R3 can be decomposed as the direct sum of S, Eκ (L) and Hc (L), i.e., (1.16) L 2 (R3 ) = S ⊕ M = S ⊕ Eκ (L) ⊕ Hc (L) , M := Eκ (L) ⊕ Hc (L) . The decomposition is not orthogonal; the three real subspaces S, Eκ (L), and Hc (L) are invariant under L but not under multiplication by i. The space M can also be characterized by (1.17) M = M E = f + ig ∈ L 2 : f, g real, f ⊥ Q E ; g ⊥ R E . From this lemma, for a fixed energy E we can decompose a wave function via (1.16). It is more convenient for our analysis to decompose the wave function into the following form: (1.18) ψ(t, x) = [Q E (x) + a E (t)R E (x) + h E (t, x)] ei2(t) with h E (t, x) ∈ M E . We shall prove in Lemma 2.2 the existence and uniqueness of such a decomposition for ψ(t, x) near a ground state. If we allow the energy E to vary, it is possible to choose E such that the component along the R E direction vanishes; i.e., there exists E(t) such that (1.19) ψ(t, x) = [Q E(t) (x) + h(t, x)]ei2(t) with h E (t, x) ∈ M E ; see Lemma 2.3. We can view this ground state Q E(t) as the best approximation to the wave function ψ(t). This choice is different from choosing the best approximation by minimizing the difference in the L 2 norm inf E,2 kψ(t, ·) − Q E ei2 k L 2 . Although (1.19) can be viewed as the best decomposition of a wave function, it introduces a time-dependent linearized operator. So for analytical purposes, it is more convenient to work with (1.18). Notice that Q E + a E (t)R E is the first-order 158 T.-P. TSAI AND H.-T. YAU approximation of Q (E+a E (t)) . So the decomposition (1.18) can be viewed as the best time-independent decomposition. Since h E (t, x) ∈ M E , we can write it as (1.20) h E (t) = ζ E (t) + η E (t) , ζ E ∈ Eκ (L) , η E ∈ Hc (L) . Theorem 1.1 gives an upper bound to the difference between the wave function and the asymptotic ground state. In fact, we can estimate the component along the excited states and the continuum spectrum more precisely in the following theorem: T HEOREM 1.3 Suppose the assumptions of Theorem 1.1 hold. Let 1/2 (1.21) ε ≡ n −1 kζ E0 ,0 k + kη E0 ,0 kY ≤ ε0 . If we decompose ψ(t) as in (1.18) with Q = Q E∞ being the profile at time t = ∞ and h(t) = ζ (t) + η(t) as in (1.20), then we have (1.22) kζ (t)k L 2 ≤ C {t}−1/2 , kη(t)k L 4 ≤ C {t}−3/4+σ , |a(t)| ≤ C {t}−1 , kη(t)k L 2 ≤ C {t}−1 , loc where {t} = {t}ε and σ = 0.01. Theorem 1.3 provides rather precise upper bounds on the asymptotic evolution. These bounds are optimal for a large class of initial data described by the next theorem. In the following, for two functions f and g, we denote (1.23) f ≈g if C1 kgk ≤ k f k ≤ C2 kgk for some constants C1 , C2 > 0. T HEOREM 1.4 (Resonance-Dominated Solutions) Assume that assumptions A0 through A2 on V hold. Suppose that the initial data ψ0 is decomposed as in (1.19) with respect to the unique E 0 , i.e., ψ0 = [Q E0 + ζ E0 ,0 + η E0 ,0 ]ei2 E0 ,0 . Write ζ E0 ,0 = z 0 u + i z 1 v and denote z E0 ,0 = z 0 + i z 1 . Suppose that these components satisfy kz E0 ,0 k kψ0 k = n < n 0 , 0 < ε := (1.24) ≤ ε0 , kη E0 ,0 kY ≤ ε2 n 2 , n where n 0 and ε0 are the same constants as in Theorem 1.1. Then the conclusions of Theorem 1.1 and Theorem 1.3 hold. In particular, there is a limit frequency E ∞ such that the estimates (1.22) hold for {t} = {t}ε . If ζ (t) = ζ E∞ (t) denotes the excited state component with respect to the linearized operator with energy E ∞ , the lower bound holds as well, and we have (1.25) ζ E∞ (t) ≈ {t}−1/2 . ε Furthermore, for all λ > 0 or λ < 0, we have kQ E∞ k2 > kQ E0 k2 and 1 (1.26) kQ E∞ k2L 2 = kQ E0 k2L 2 + kζ E0 ,0 k2L 2 + o(ε2 n 2 ) . 2 DYNAMICS OF NLS 159 Solutions satisfying (1.25) are called resonance-dominated solutions. Roughly speaking, Theorem 1.4 states that, if the excited state component is much bigger than the dispersive component, then the decay mechanism is dominated by resonance. Furthermore, (1.26) states that approximately half of the probability density of the excited state, |ζ E0 (x)|2 /2, is transferred to the ground state independently of the sign of λ. Since the total probability is conserved, the other half is transferred to the dispersive part. We have decomposed the wave function according to the optimal energy E 0 (1.19). Alternatively, we can first fix an energy E ∗ and decompose the initial data ψ0 as (1.27) ψ0 = [Q E∗ + a∗ R E∗ + ζ∗ + η∗ ]ei2∗ . If the components satisfy (1.28) kψ0 k = n < n 0 , kζ∗ k = εn , kη∗ kY ≤ Cε2 n 2 , 0 < ε ≤ ε0 , |a∗ | ≤ ε2 n 2 , then we can re-decompose ψ0 as in Theorem 1.4 with respect to E 0 satisfying the estimate (1.24); see Lemma 2.3. This implies that the set of all such ψ0 contains an open set with nonlinear ground states in its boundary. Therefore, the class of resonance-dominated solutions is in a sense large. The condition (1.24) is very subtle. It states that the excited-state component is much bigger than the dispersive component with respect to a decomposition according to L. Since L differs from HE by order n 2 , the condition (1.24) does not hold for a decomposition of ψ0 with respect to the linear Hamiltonian HE . Finally, the following existence result shows that the resonance-dominated solutions are not all solutions. We expect that the dispersion-dominated solutions constructed by the following theorem are rare: T HEOREM 1.5 (Dispersion-Dominated Solutions) Let assumptions A0 through A2 apply on V . Let Z = H 2 ∩W 2,1 (R3 ). For a given nonlinear ground state Q E∞ with kQ E∞ k = n ≤ n 0 , let E = E ∞ and L = L E∞ . For any given ξ∞ ∈ Hc (L) ∩ Z with sufficiently small Z norm, there exists a solution ψ(t) of (1.1) and a real function θ (t) = O(t −1 ) for t > 0 so that kψ(t) − ψas (t)k H 2 (R3 ) ≤ Ct −2 , where ψas (t) = Q E e−i Et+iθ (t) + e−i Et et L ξ∞ . In particular, ψ(t) − Q E∞ e−i Et+iθ (t) = O(t −3/2 ) in L 2loc . Theorem 1.5 constructs dispersion-dominated solutions based on the operator et L . The scattering property of this operator is very similar to the standard operator ei1t . In particular, for χ∞ ∈ e Z = H 2 (R3 ) ∩ W 2,1 (R3 , (1 + |x|4 )d x) sufficiently 160 T.-P. TSAI AND H.-T. YAU small in e Z with χ̂∞ (0) = 0 and ∇ χ̂∞ (0) = 0, the same statement holds if we replace ψas by eas (t) = Q E e−i Et + ei1t χ∞ . ψ This will be shown in the proof of Theorem 1.5. The resonance solutions were first observed by Buslaev and Perel0 man [3] for some one-dimensional Schrödinger equation with two bound states and a higher nonlinear term (we thank the referee for supplying us with this reference). In the physical dimension d = 3, this type of solution was proven by Soffer and Weinstein in an important paper [16], but for real solutions to the nonlinear Klein-Gordon equation (1.29) ∂t2 u + B 2 u = λu 3 , B 2 := (−1 + V + m 2 ) , where λ is a small nonzero number. Assume that B 2 has only one eigenvector (the ground state) φ, B 2 φ = 2 φ, with the resonance condition < m < 3 (and some positivity assumption similar to that appearing in assumption A1). Rewrite real solutions to equation (1.29) as u = aφ + η with a(t) = Re A(t)eit , Re A0 (t)eit = 0 . Then A and η satisfy the equations (1.30) (1.31) 1 −it e φ, λ(aφ + η)3 2i 2 2 (∂t + B )η = Pc λ(aφ + η)3 . Ȧ = Theorem 1.1 in [16] states that all solutions decay as A(t) ∼ hti−1/4 , kη(t)k L ∞ ∼ hti−3/4 . In particular, the ground state is unstable and will decay as a resonance with rate t −1/4 . We first remark that the proof in [16] has only established the upper bound t −1/4 . Furthermore, a universal lower bound of the form t −1/4 is in fact incorrect. From the previous work of [1, 5], it is clear that dispersion-dominated solutions decaying much faster than t −1/4 exist. Similar to Theorems 1.4 and 1.5, we have the following two cases: (1) η(0) A(0): The dominant term on the right side of (1.30) is λa 3 φ 3 . (2) η(0) A(0): The dominant term is λη3 . In case 2 another type of solutions arises, namely, those with decay rate A(t) ∼ hti−2 , kηk L ∞ ∼ hti−3/2 . We shall sketch a construction of such solutions at the end of Section 8. Notice that all solutions in [16] decay as a function of t. Therefore, we can view [16] as a study of asymptotic dynamics around a vacuum. Although most works concerning asymptotic dynamics of nonlinear evolution equations have been concentrated on cases with vacuum as the unique profile at t = ∞, more interesting DYNAMICS OF NLS 161 and relevant cases are asymptotic dynamics around solitons (such as the Hartree equations [5]; see also [2, 3]). The soliton dynamics have extra complications involving translational invariance. The current setting of nonlinear Schrödinger equations with local potentials eliminates the translational invariance and constitutes a useful intermediate step. This greatly simplifies the analysis but preserves a key difficulty that we now explain. Recall that we need to approximate the wave function ψ(t) by nonlinear ground states for all t. Since we aim to show that the error between them decay like t −1/2 , we have to track the nonlinear ground states with accuracy at least like t −1/2 . Although the nonlinear ground states approximating the wave function ψ(t) can in principle be defined, say, via equation (1.4) or (1.6), neither characterization is useful unless we know the wave functions ψ(t) precisely. Furthermore, even assuming we can track the approximate ground states reasonably well, the linearized evolution around these approximate ground states will be based on time-dependent, non-self-adjoint operators Lt . At this point we would like to mention the approach of [15] based on perturbation around the unitary evolution eit H1 , where H1 is the original self-adjoint Hamiltonian. While we do not know whether this approach can be extended to the current setting by adding the ideas of [16] (it was announced in [16] that its method can be extended to (1.1) as well), such an approach can be difficult to extend to the Hartree or other equations with nonvanishing solitons. The main reason is that these dynamics are not perturbations of linear dynamics. We believe that perturbation around the profile at t = ∞ is a more natural setup. In this approach, at least we do not have to worry about the time dependence of approximate ground states in the beginning. But the linearized operator L = L∞ is still non-self-adjoint, and it does not commute with the multiplication by i. So calculations and estimates based on L are rather complicated. Our idea is to map this operator to a self-adjoint operator by a bounded transformation in Sobolev spaces. This map in a sense brings the problem back to the self-adjoint case for various calculations and estimates. Another important input we used is the existence and boundedness of the wave operator for L. The boundedness of the wave operator in Sobolev spaces for the standard one-body Schrödinger operator is a classical theorem of Yajima [20]. In the current setting it was recently proved by Cuccagna [4]. See Section 2 for more details. The next step is to identify and calculate the leading oscillatory terms of the nonlinear systems involving the bound-states components and the continuum-spectrum components. The leading-order terms, however, depend on the relative sizes of these components, and thus we have two different asymptotic behaviors: The resonance-dominated solutions and the dispersion-dominated solutions. Finally, we represent the continuum spectrum component in terms of the bound-states components, and this leads to a system of ordinary differential-integral equations for the bound-states components. This system can be put into a normal form, and the size of the excited-state component can be seen to decay as t −1/2 . Notice that the 162 T.-P. TSAI AND H.-T. YAU phase and the size of the excited-state component decay differently. It is thus important to isolate the contribution of the phase in the system. Finally, we estimate the error terms using estimates of the linearized operators. These estimates are based on standard methods from scattering theory [6, 14, 16] and estimates on the wave operators [4, 20]. 2 Preliminaries We first fix our notation. Let H k denote the Sobolev spaces W k,2 (R3 ). The weighted Sobolev space L r2 (R3 ) is defined in (1.7). The inner product ( , ) is Z (2.1) ( f, g) = f¯g d 3 x . R3 For two functions f and g, we denote f = O(g) if k f k ≤ C kgk for some constant C > 0 . We will denote, as in (1.23), f ≈g if C1 kgk ≤ k f k ≤ C2 kgk for some constants C1 , C2 > 0 . For a double index α = (α0 , α1 ), α0 and α1 nonnegative integers, we denote (2.2) z α = z α0 z̄ α1 , (32) |α| = α0 + α1 , [α] = −α0 + α1 . = z z̄ . We will define z = µ−1 p where µ = eiκt . Hence For example, z z α = µ[α] p α . In what follows |α| = 2 and |β| = 3. 3 2 2.1 Nonlinear Ground States Family We establish here the existence and basic properties of the nonlinear groundstates family mentioned in Section 1. Our statement and proof are also valid for nonlinear excited states. L EMMA 2.1 Suppose that −1 + V satisfies assumptions A0 and A2. Then there is an n 0 sufficiently small such that for E between e0 and e0 +λn 20 there is a nonlinear ground state {Q E } E solving (1.2). The nonlinear ground state Q E is real, local, and smooth, λ−1 (E − e0 ) > 0, and Z −1/2 Q E = nφ0 + O(n 3 ) , n ≈ C[λ−1 (E − e0 )]1/2 , C = φ04 d x . Moreover, we have R E = ∂ E Q E = O(n −2 )Q E + O(n) = O(n −1 ) and ∂ E2 Q E = O(n −3 ). If we define c1 ≡ (Q, R)−1 , then c1 = O(1) and λc1 > 0. P ROOF : Suppose that Q = Q E is a nonlinear ground state satisfying (−1 + V − e0 )Q + λQ 3 = E 0 Q where E 0 = E − e0 is small. Write Q = nφ0 + h with real h ⊥ φ0 . Then h satisfies (2.3) (−1 + V − e0 − E 0 )h + λ(nφ0 + h)3 = E 0 nφ0 . DYNAMICS OF NLS 163 Taking the inner product with φ0 , we see that it is necessary that λ and E 0 be of the same sign. For E 0 sufficiently small, since the spectral gap of the operator −1 + V − e0 is of order 1, the same conclusion holds for the operator −1 + V − e0 − E 0 . Thus we can check directly from (2.3) that (2.4) kQ E k := n 2 ∼ λ−1 E 0 , khk2 = O(λn 3 ) . The existence of solutions h and E 0 can be established from the implicit function theorem, or the solutions can simply be obtained by a contraction mapping argument. By differentiating the equation of Q E with respect to E, we have (2.5) L + RE = Q E , L + := −1 + V − E + 3λQ 2 . Denote by |0i the ground state to L + with L 2 norm 1 and eigenvalue ≈ λn 2 . We have Q E = O(n)|0i + O(λn 3 ) . Hence (using the spectral gap) n |0i + O(λn 3 ) R = (L + ) [O(n)|0i + O(λn )] = O λn 2 −1 (2.6) 3 = O(λ−1 n −2 )Q + O(n) . Hence c1 = O(1). Since the sign of the ground-state energy to L + is the same as λ, we have λ(Q, R) > 0. By differentiating the equation, we can prove ∂ E2 Q E = O(n −3 ) in a similar way. Suppose Q E = O(n) and a is a small parameter with |a| n. By Taylor expansion and ∂ E2 Q E = O(n −3 ), we have (2.7) Q E+a = Q E + a R E + O(n −3 a 2 ) = Q E + O(n −1 a) . Since (a R E )2 = O(n −2 a 2 ), from (2.7) we also have (2.8) kQ E+a k2 = kQ E k2 + 2a(Q E , R E ) + O(n −2 a 2 ) . We will need the following renormalization lemmas. We will formulate them in terms of nonlinear ground states, although they also hold for nonlinear excited states. Let Y1 be either Y , defined in (1.8), or L 2−r0 , defined in (1.7) with r0 > 3. Recall from Lemma 1.2 that h ∈ M if and only if Re h ⊥ Q and Im h ⊥ R. L EMMA 2.2 Let ψ ∈ Y1 be given with kψkY1 = n, 0 < n ≤ n 0 . Suppose that for some nonlinear ground state Q E ei2 we have kψ − Q E ei2 kY1 ≤ Cτ n, 0 ≤ τ ≤ ε0 . Then there are unique small a, θ, and h such that (2.9) ψ = [Q E + a R E + h] ei(2+θ ) , Furthermore, kQ E kY1 > 9 n, 10 h ∈ ME . a = O(τ n 2 ), θ = O(τ ), and h = O(τ n). 164 T.-P. TSAI AND H.-T. YAU P ROOF : By considering ψe−i2 instead of ψ, we may assume 2 = 0. Write ψ = Q + k with a small k = O(τ n). We want to solve small a and θ such that h = ψe−iθ − Q E − a R E ∈ M E , i.e., 0 (Q, Re(Q + k)e−iθ − Q − a R) = , 0 (R, Im[(Q + k)e−iθ ]) with the left side of order [ τ τn 2 ] if (a, θ ) = (0, 0). We compute the derivative of this vector with respect to a and θ at (a, θ ) = (0, 0) and obtain the matrix −(Q, R) (Q, Im k) −(Q, R) O(τ n 2 ) , = 0 (R, − Re(Q + k)) 0 −(R, Q) + O(τ ) which is invertible. Here we have used Q = O(n) and R = O(n −1 ). Since (Q, R) = O(1), by the implicit function theorem we can solve a = O(τ n 2 ) and θ = O(τ ) satisfying the equation. Hence h = (Q E + k)e−iθ − Q E − a R E = O(τ n). The next lemma shows that the best decomposition as defined by (1.19) can always be achieved for a small vector near a nonlinear ground state. L EMMA 2.3 Let ψ ∈ Y1 with kψkY1 = n bounded by a small number n 0 , i.e., 0 < n ≤ n 0 . If ψ is near a nonlinear ground state Q E1 ei21 in the sense that kψ − Q E1 ei21 kY1 ≤ Cτ n for some τ with 0 ≤ τ ≤ ε0 and 0 small. Then there is a unique E near E 1 such that the component along the R E direction as defined by (2.9) vanishes; i.e., there is unique small θ and h such that ψ = [Q E + h] ei(21 +θ ) , h ∈ ME . Moreover, E − E 1 = O(τ n ), θ = O(τ ), and h = O(τ n). If ψ is given explicitly as ψ = [Q E1 + a1 R E1 + h 1 ]ei21 with h 1 ∈ M E1 and a1 = O(τ 2 n 2 ), we have 5 (2.10) |E − E 1 | ≤ |a1 | = O(τ 2 n 2 ) , θ = O(τ 3 ) , h − h 1 = O(τ 2 n) . 4 2 2 P ROOF : As in the proof of the previous lemma, we may assume 21 = 0. Write ψ = Q E1 + k1 with h 1 ∈ M E1 and k1 = a1 R E1 + h 1 = O(τ n) small. Let E = E 1 + γ . We want to find small γ and θ such that h = ψe−iθ − Q E ∈ M E , that is, (Q E1 +γ , Re(Q E1 + k1 )e−iθ − Q E1 +γ ) 0 (2.11) = . 0 (R E1 +γ , Im[(Q E1 + k1 )e−iθ ]) When (γ , θ ) = (0, 0), the left side equals 2 (Q E1 , a1 R E1 ) τn , which is of order . 0 0 The derivative of this vector with respect to γ and θ at (γ , θ) = (0, 0) is given by DYNAMICS OF NLS 165 (Q E1 , Im k1 ) (R E1 , −Q E1 + Re k1 ) = (∂ E R E | E=E1 , Im k1 ) (R E1 , − Re(Q E1 + k1 )) −(Q, R) O(τ n 2 ) , O(τ n −2 ) −(R, Q) + O(τ ) which is invertible. Here we have used Q = O(n), R = O(n −1 ), and ∂ E R E = O(n −3 ). By the implicit function theorem we can solve γ = O(τ n 2 ) and θ = O(τ 2 ) satisfying the equation. Hence h = (Q E1 + k1 )e−iθ − Q E1 +γ = O(τ n). Finally, when a1 = O(τ 2 n 2 ), the vector on the left side of (2.11) at (γ , θ) = (0, 0) is of order 2 2 τ n . 0 Hence we can solve γ = O(τ 2 n 2 ) and θ = O(τ 3 ). We also have h − h 1 = (Q E1 + a1 R E1 + h 1 )e−iθ − h 1 − Q E1 +γ = O(nθ) + O(n −1 a1 ) = O(τ 2 n). The last statement is proven. The next lemma provides estimates on the components in (1.18) when the reference nonlinear ground state varies. L EMMA 2.4 Let ψ = [Q E1 + a1 R E1 + h 1 ]ei21 with kQ E1 kY1 = n ≤ n 1 , kh 1 kY1 = ρ ≤ ε1 n, and |a1 | ≤ Cρ 2 . Suppose E 2 = E 1 + γ with |γ | ≤ Cρ 2 . By Lemma 2.2 we can rewrite ψ uniquely with respect to E 2 as (2.12) ψ = Q E2 + a2 R E2 + h 2 ei22 , where h 2 ∈ M E2 , a2 , and θ = 22 − 21 are small. We have the estimates θ = O(n −3 ργ ) , E 1 + a1 − E 2 − a2 = O(n −1 ργ ) , (2.13) h 1 − h 2 = O(n −2 ργ ) . Notice that n −3 ργ ≤ Cε13 is small. P ROOF : Note |γ | ≤ Cρ 2 ≤ ε12 n 2 . Using (2.7) we have Q E1 − Q E2 = O(n −1 γ ) , (2.14) R E1 − R E2 = O(n −3 γ ) , λQ 2E1 − λQ 2E2 = O(γ ) . Since e−i21 ψ − Q E2 = [Q E1 − Q E2 ] + a1 R E1 + h 1 = O(n −1 γ ) + O(ρ 2 n −1 ) + O(ρ) = O(ρ) ≤ Cε1 n , Lemma 2.2 is applicable with τ = ρ/n, and we have (2.12) with a2 = O(τ n 2 ) = O(nρ) , h 2 = O(τ n) = O(ρ) , θ = O(τ 2 ) = O(n −2 ρ 2 ) . We have (2.15) [Q E1 + a1 R E1 + h 1 ]e−iθ = Q E2 + a2 R E2 + h 2 . 166 T.-P. TSAI AND H.-T. YAU Denote by (RS) the right side and by (LS) the left side. Taking the imaginary part and then taking the inner product with R E2 , we have (R E2 , Im(RS)) = (R E2 , Im h 2 ) = 0 , since h ∈ M E2 , and (R E2 , Im(LS)) = (R E2 , Q E1 + a1 R E1 + Re h 1 )(− sin θ) + (R E2 , Im h 1 ) cos θ . Since (R E2 , Q E1 + a1 R E1 + Re h 1 ) ≈ 1 and (R E2 , Im h 1 ) = (R E1 + O(n −3 γ ), Im h 1 ) = 0 + O(n −3 ργ ) , we conclude that sin θ +O(n −3 ργ ) = 0. Since θ is small, we have θ = O(n −3 ργ ). Write eiθ = 1 + O(θ ) in (2.15). Since [Q E1 + a R E1 + h 1 ]O(θ) = O(n)O(n −3 ργ ) , equation (2.15) gives (2.16) Q E1 + a1 R E1 − Q E2 − a2 R E2 = O(n −2 ργ ) + h 2 − h 1 . Taking the real part and then taking the inner product with Q E2 , we get from the right side O(n · n −2 ργ ) + 0 + (Q E1 + O(n −1 γ ), h 1 ) = O(n −1 ργ ). On the other hand, by (2.7) Q E 1 + a1 R E 1 − Q E 2 − a2 R E 2 = Q E2 − γ R E2 + O(γ 2 n −3 ) + a1 R E2 + O(γ n −3 ) − Q E2 − a2 R E2 = (−γ + a1 − a2 )R E2 + O(n −3 ρ 2 γ ) , where we have used E 1 = E 2 − γ . Hence we conclude (E 1 − E 2 + a1 − a2 )(Q E2 , R E2 ) + O(n −3 ρ 2 γ ) = O(n −1 ργ ) . Since (Q E2 , R E2 ) = O(1), we have E 1 − E 2 + a1 − a2 = O(n −1 ργ ). From (2.16), we thus have h 2 − h 1 = O(n −1 ργ )R E2 + O(n −3 ρ 2 γ ) − O(n −2 ργ ) = O(n −2 ργ ) . 2.2 Linearized Operator: Spectral Analysis We now study the spectral properties of L. In this section, we identify C with R and the multiplication by i becomes 0 −1 J= . 1 0 2 Recall the definition of M and the decomposition L 2 (R3 ) = S ⊕ M. This decomposition is nonorthogonal and presents some problems in analysis. Let X be the space orthogonal to Q, ⊥ Q 2 2 3 X = 5(L ) = φ ∈ L (R ) : φ ⊥ Q ; X ←→ , Q⊥ DYNAMICS OF NLS 167 where 5 is the orthogonal projection that eliminates the Q direction: (Q, h) Q. (Q, Q) We claim that there is a “nice” operator U from M to X and a self-adjoint operator A that is a perturbation of H1 such that (2.17) L = −U −1 J AU . 5h = h − M Let PM be the projection from L 2 onto M according to the decomposition L (R3 ) = S ⊕ M. It has the matrix form 2 ⊥ L P1 0 Q PM : −→ , PM = , 0 P2 R⊥ L2 2 where the projections are given by (2.18) P1 : L 2 −→ Q ⊥ , P1 = Id − c1 |RihQ| , P2 : L 2 −→ R ⊥ , P2 = Id − c1 |QihR| . c1 = (Q, R)−1 , Clearly P1 R = 0 and P2 Q = 0. One can easily check that I 0 I 0 : X −→ M . (2.19) RM X ≡ : M −→ X , R X M ≡ 0 5 0 P2 We now define (2.20) 1/2 1/2 1/2 A := (H 2 + H 1/2 2λQ 2 H 1/2 ) = H L + H 1/2 , H = L− . A is a self-adjoint operator acting on L 2 (R3 ), with Q as an eigenvector with eigenvalue 0. We shall often view A as an operator restricted to its invariant subspace X . Define 1/2 −1/2 A H 0 (2.21) U0 : X −→ X , U0 ≡ , 0 A−1/2 H 1/2 and let (2.22) U = U0 R M X : M −→ X , U −1 = R X M U0−1 : X −→ M . Notice that H −1/2 is defined only on Q ⊥ . We can easily check that U −1U | M = Id M and UU −1 | X = Id X . Moreover, we have 0 H 1/2 A−1/2 A3/2 H −1/2 0 −1 −(U J )(AU ) M = 0 A1/2 H 1/2 −P2 H −1/2 A1/2 0 0 H = . −P2 5L + 0 Since P2 5L + = L + when acting on R ⊥ , we have proven (2.17). If we define 1 (2.23) U± = 5A1/2 H −1/2 P1 ± 5A−1/2 H 1/2 5 , 2 168 T.-P. TSAI AND H.-T. YAU then (U± )∗ = 12 (P2 H −1/2 A1/2 5 ± 5H 1/2 A−1/2 5) and (2.24) U = U+ + CU− , U −1 = (U+ )∗ − C(U− )∗ , where C = [ 10 −10 ] is the conjugation operator in L 2 (R3 ). Although U+ and U− are not self-adjoint operators, they commute with the multiplication operator i. Hence, as operators in L 2 (R3 ), (2.25) U i = (U+ + CU− )i = i(U+ − CU− ) . Since the mass of Q E is small, A is a self-adjoint perturbation of H1 = −1 + V −e0 . As H1 has two simple eigenvalues, so does A. The ground state of A is just Q (which with the normalized ground state equals φ0A = Q/ kQk2 ) with energy 0. Standard perturbation theory implies that A has an excited state Aφ1A = κφ1A such that κ = e10 + O(n 2 ) and φ1A = φ1 + O(n 2 ). The spectral decomposition Lemma 1.1 is a corollary of (2.17) and the spectral properties of A. For example, define A u −1 φ1 . =U v φ1A We can check easily that u and v are the generalized eigenvectors of (1.13). We summarize the results here in the following lemma: L EMMA 2.5 Let X be the orthogonal complement of Q in L 2 . Let A be defined by (2.20). Let U and U −1 be defined by (2.22). Then (2.17) holds, i.e., L| M = −U −1 J AU . If we denote by PκA and PcA the orthogonal projections onto the eigenspace and continuous spectrum space of A, we have (2.26) U PM = U , U PκL = PκA U , A U PL c = Pc U . Furthermore, we have U = U+ + U− C and U −1 = U+∗ − U−∗ C, with U+ and U− commuting with i. 2.3 The Equations In this subsection we derive the equations for the components of the solution ψ(t) of the Schrödiger equation (1.1) according to the decomposition (1.18). Fix an energy E, the corresponding ground state Q = Q E , and its tangent R = R E = ∂ E Q E . Recall the decomposition (1.18) of the wave function ψ(t, x). We let (2.27) 2(t) = θ(t) − Et . Denote by h a = a(t)R E (x) + h(t, x). Substituting the above ansatz into (1.1), we have the equations ∂t h a = Lh a + i −1 (F + θ̇ (Q + h a )) , F = λQ(2|h a |2 + h a2 ) + λ|h a |2 h a = λQ(2|h|2 + h 2 ) + 2λQ Ra(2h + h̄) + 3λQ R 2 a 2 + λ(a R + h)2 (a R + h̄) , DYNAMICS OF NLS 169 where L is already defined in (1.10). By (1.2) and (2.5), we have (L − ∂t )a R = −ia Q − ȧ R. Hence ∂t h = Lh + i −1 (F + θ̇(Q + a R + h)) − ia Q − ȧ R . Since h(t) ∈ M and M is an invariant subspace of L, we have i −1 (F + θ̇(Q +a R + h)) − ia Q − ȧ R ∈ M. From the characterization M = [ f + ig : f ⊥ Q, g ⊥ R], the functions a(t) and θ (t) satisfy (Q, Im(F + θ̇ h) − ȧ R) = 0 , (R, Re(F + θ̇(Q + a R + h)) + a Q) = 0 . The equation on M is (2.28) ∂t h = Lh + PM Fall , Fall = i −1 (F + θ̇(a R + h)) . Recall the representation of L in terms of A. The previous equation can be rewritten as ∂t (U h) = −i A(U h) + U PM Fall . (2.29) We can write U h = of A. Then we have zφ1A + w where w denotes the part on the continuum spectrum ż = −iκz + (φ1A , U PM Fall ) , ẇ = −i Aw + PcA U PM Fall . A A A Since U PL c = Pc U , we have Pc U PM = Pc U . Using the definitions of u and v, we have (φ1A , U PM Fall ) = (v, Re Fall ) + i(u, Im Fall ) = (u + , Fall ) − (u − , F all ) where u ± = 12 (u ± v). If we define p(t) = eiκt z(t), we have e−iκt ṗ(t) = i −1 {(u + , F) + (u − , F) + [(u + , h) + (u − , h) + (u, R)a]θ̇} . Summarizing, we have the decomposition ψ = (Q + a R + h)e−i Et+iθ , U h = zφ1A + w , z = e−iκt p , µ = eiκt , and the equation −1 ȧ = (c1 Q, Im(F + θ̇ h)) , c1 = (Q, R) , (2.30) ∂t w = −i Aw + PcA U i −1 (F + θ̇(a R + h)) , −iκt ṗ = (u + , F) + (u − , F) + [(u + , h) + (u − , h) + (u, R)a]θ̇ , ie where (2.31) F = λQ(2|h|2 + h 2 ) + 2λQ Ra(2h + h̄) + 3λQ R 2 a 2 + λ(a R + h)2 (a R + h̄) , (2.32) θ̇ = − [a + (c1 R, Re F)] · [1 + a(c1 R, R) + (c1 R, Re h)]−1 . This is a system of equations involving a, z, and η only. It is the system that we shall solve in the rest of this paper. Note that θ does not appear explicitly in the system (2.30). We see θ̇ only on the right side. Although θ will appear when we 170 T.-P. TSAI AND H.-T. YAU integrate the main term of η, it appears only in the form eiθ ; hence we do not need estimates for θ. We also have the following decomposition of h according to (1.16): η = U −1 w ∈ Hc (L) , (2.33) h = ζ +η, (2.34) ζ = U −1 zφ1A = Re(z)u + Im(z)iv = zu + + z̄u − ∈ Eκ (L) . The equation for η can be obtained from that of w or more directly from projecting the original equation of h onto the continuum spectrum of L: (2.35) −1 ∂t η = Lη + PL c i (F + θ̇(a R + h)) . 2.4 Linear Estimates Here we collect a few estimates on the operators U , L, and A. They will be proven at the end of this paper. We denote the wave operators for L (respectively, A and H ) by WL (respectively, W A and W H ). They are defined by WL = lim e−t L e−it H∗ , (2.36) t→∞ W H = lim e t→∞ W A = lim eit A e−it H∗ , t→∞ it H −it H∗ e , where (2.37) H∗ = −1 − E . Note that if W A exists, we have the intertwining property that f (A)Pc (A) = W A f (H∗ )W A∗ for suitable functions f . We also have a similar property for L. L EMMA 2.6 For k = 0, 1, 2, there is a positive constant C1 so that (2.38) C1−1 kφk H k ≤ ke−it A φk H k ≤ C1 kφk H k for all φ ∈ X ∩ H k and all t ∈ R. L EMMA 2.7 (Decay Estimates for e−it A ) For q ∈ [2, ∞] and q 0 = q/(q − 1), (2.39) −it A A −3 12 − q1 e kφk L q 0 . Pc 5φ L q ≤ C|t| For smooth local functions φ and sufficiently large r1 , we have −r −it A 1 A −r −9/8 1 hxi 1 e (2.40) P 5hxi φ c 2 ≤ Chti (A − 0i − 2κ)l L where l = 1, 2 and 0i means σ i with limσ →0+ outside of the bracket. Estimate (2.39) for A = −1 +√V was proven in [8] using estimates from [7] and [10]. Estimate (2.40) for A = H was proven by Soffer and Weinstein [16]. DYNAMICS OF NLS L EMMA 2.8 (2.41) φ0 φ12 , Im 171 1 A 2 P 5φ0 φ1 = A − 0i − 2κ c 1 φ0 φ12 , Im PcH1 φ0 φ12 + O(n 2 ) > 0 . H1 − 0i − 2κ This lemma is a perturbation result. Notice that κ = e01 + O(n 2 ); hence the second term is greater than γ0 by assumption A1. (i) The operators U0 and U0−1 are bounded operaL EMMA 2.9 (Operator U ) tors in W k, p ∩ X for k ≤ 2, 1 ≤ p < ∞, and in L r2 ∩ X for |r | ≤ r0 . (L r2 is the weighted L 2 space defined in (1.7).) Hence U : M → X and U −1 : X → M are bounded in W k, p and L r2 norms. (ii) The commutator [U, i] is a local operator in the sense that (2.42) k[U, i]φk L 8/7 ∩L 4/3 ≤ C kφk L 4 . L EMMA 2.10 (Wave Operators) The wave operators WL and W A defined by (2.36) exist and satisfy W k, p estimates for k ≤ 2, 1 ≤ p < ∞ (similar estimates hold for their adjoints): WL PL k, p k, p ≤ C , W A P A 5 k, p k, p ≤ C . c (W ,W ) c (W ,W ) The statement on WL was proven in [4], following the proof of [20]. Hence we need only to prove the statement on W A . 3 Main Oscillation Terms We now identify the main oscillation terms in equation (2.30). Most quantities treated in this paper, such as F, a, z, and η, are strongly oscillatory. Hence it is necessary to identify their oscillatory parts before we can estimate. We shall use the complex amplitude of the excited state, z, as the reference. Recall z(t) = e−iκt p(t). We will show that p(t) ∼ t −1/2 and its oscillation (phase speed) is much smaller than κ. The change of mass on the direction of the nonlinear ground state is given by a. We will also show that a = O(z 2 ) and the order of the dispersive wave η(t) is also of order O(z 2 ). Assuming these orders, the second-order term in F is given explicitly by (3.1) F (2) = λQ(2|ζ |2 + ζ 2 ) = z 2 φ(20) + z z̄φ(11) + z̄ 2 φ(02) where ζ = zu + + z̄u − and φ(20) = λQ(u 2+ + 2u + u − ) = λQφ12 + O(n 3 ) , (3.2) φ(11) = 2λQ(u 2+ + u 2− + u + u − ) = 2λQφ12 + O(n 3 ) , φ(02) = λQ(u 2− + 2u + u − ) = O(n 3 ) . 172 T.-P. TSAI AND H.-T. YAU We shall write F (2) = z α φα , where α is a double indices (i j) with i + j = 2 and i, j ≥ 0. The repeated indices mean summation. We shall use β later on to denote double indices summing to 3 and γ summing to 4. 3.1 Leading Oscillatory Terms in a We first identify the main oscillation terms of a(t). Recall from (2.30) that ȧ = (c1 Q, Im F+θ̇ h) and c1 = (Q, R)−1 . We shall impose the boundary condition of a at t = T . This is in fact a condition imposed on the choice of E T . Hence we use the following equivalent integral equation: Z t (3.3) a(t) = a(T ) + (c1 Q, Im F + θ̇ h)(s)ds . T (2) The main term of Im F + θ̇ h is ImR F = Im λQζ 2 , and thus the leading oscillation t term of a(t) is from the integral T A(2) ds with A(2) = (c1 Q, λQ Im ζ 2 ) . Since ζ = zu + + z̄u − , we have Im ζ = Im z(u + − u − ) and Im ζ 2 = (Im z 2 )(u 2+ − u 2− ); therefore, we have A(2) = C1 Im z 2 , C1 = (c1 Q, λQ(u 2+ − u 2− )) . Write z 2 (s) = e−2iκs p 2 = (3.4) 1 d −2iκs 2 )p . (e −2iκ ds We can integrate A(2) by parts to get Z t Z t Z t (2) 2 2 t −2iκs (3.5) A ds = Im C1 z ds = 2a20 Re [z ]T − e 2 p ṗ ds , T T T where C1 1 c1 Q, λQ(u 2+ − u 2− ) = O(n 2 ) = 4κ 4κ and the last integral is a higher-order term. Let us denote (3.6) a20 = a (2) (t) = a20 (z 2 + z̄ 2 )(t) , (3.7) which is the main oscillatory term in a. We denote the rest of a by b, i.e., a(t) = a (2) (t) + b(t) , (3.8) and we have b(t) = a(T ) − a (2) (T ) Z t (3.9) + (c1 Q, Im[F − F (2) + θ̇ h]) − 4a20 Re e−2iκs p ṗ (s)ds . T It is easy to see that a, b, ȧ = O(z 2 ), but ḃ = O(z 3 ). In other words, a (2) is the part of a with strong oscillation and b is the part of a that has slower oscillation. The use of b is convenient when we work on normal forms of p and a later. (We DYNAMICS OF NLS 173 will integrate by parts terms involving b, and in some sense replace terms involving b by terms involving ḃ, which is smaller. If we work with a instead, then we replace a by ȧ, which is of the same order and we need one more step.) In fact, in principle one should treat the normal forms of p and a together, since they correspond to the excited states and the ground state. See [17] for more elaboration on this point. It should be noted that, although b(t) is not oscillatory, it is in fact larger than a (2) (t), and hence is the main term of a(t). Another point to make here is that the introduction of b(t) is for computational convenience. The true variable we work with is still a(t). In particular, the induction assumption we make later in Section 5 is on a, not b. Notice that when integrating θ̇, we take Re F instead of Im F. Hence the term C|z|2 survives and cannot be integrated; thus θ(t) = O(log t), although a(t) = O(t −1 ). 3.2 Leading Oscillatory Terms in η We now identify the main oscillation term in η. From the basic equation (2.30), we rewrite the w equation as ∂t w = −i Aw − i θ̇ w + PcA U i −1 [F + θ̇(a R + ζ )] − PcA [U, i]θ̇η . where we have used the commutator [U, i] to interchange U and i so as to produce the term i θ̇ w. This term is a global linear term in η and cannot be treated as an error term (however, [U, i]θ̇η is an error term). We can eliminate it by rewriting the last equation in terms of η̃ = eiθ w = eiθ U η, i.e., ∂t η̃ = −i Aη̃ + eiθ PcA U i −1 [F + θ̇(a R + ζ )] − eiθ PcA [U, i]θ̇η , or in integral form, (3.10) η̃(t) = e−i At η̃0 + Z t e−i A(t−s) PcA Fη (s)ds , 0 where (3.11) Fη = eiθ U i −1 [F + θ̇(a R + ζ )] − eiθ [U, i]θ̇η . Since U and U −1 are bounded in Sobolev spaces by Lemma 2.9 and (3.12) η(t) = U −1 e−iθ (t) η̃(t) , for the purpose of estimation we can treat η and η̃ as the same. Let Fη,2 be the second-order term of Fη , and let Fη,3 denote the rest, i.e., Fη = Fη,2 + Fη,3 , (3.13) Fη,2 = eiθ U i −1 F (2) = eiθ U i −1 z α φα , Fη,3 = eiθ U i −1 (F − F (2) ) + θ̇(a R + ζ ) − eiθ [U, i]θ̇η . 174 T.-P. TSAI AND H.-T. YAU Since U = U+ + CU− with U+ and U− commuting with i, see (2.24), we can write Fη,2 as Fη,2 = eiθ (U+ + CU− )i −1 z α φα (3.14) 8(20) (3.15) = i −1 eiθ (U+ − CU− )z α φα = i −1 eiθ z α 8α , = U+ φ(20) − U− φ(02) , 8(02) = U+ φ(02) − U− φ(20) , 8(11) = (U+ − U− )φ(11) . Replacing Fη in (3.10) by Fη,2 and integrating by parts, we have Z t Z t e−i A(t−s) PcA Fη,2 ds = e−it A eis(A−0i+[α]κ) (eiθ p α )(s) i −1 PcA 8α ds 0 0 ηα + (∗) , = η̃(2) (t) − e−it A (eiθ z α )(0)e (3.16) where (3.17) η̃(2) (t) = eiθ(t) z α (t)e ηα , e ηα = −1 P A 8α , A − 0i + [α]κ c and (∗) denotes the remainder term from integration by parts, Z t d e−i(t−s)A eisκ[α] (eiθ p α )e ηα ds . (∗) = − ds 0 In the integration we have inserted the regularizing factor eis(−0i) with the sign of 0i chosen so that e−it Ae ηα decays as t → +∞. See Lemma 2.7. We shall see that η̃(2) (t) is the main oscillation term in η̃(t). We denote the rest by η̃(3) (t): η̃ = η̃(2) + η̃(3) . (3.18) Notice that this is not a decomposition in L 2 : Both η̃(2) and η̃(2) are not in L 2 . Nevertheless, this decomposition is useful for studying the local behavior of η̃. Also note that, although η̃α 6∈ L 2 but is still “orthogonal” to the eigenvector of A, (φ1A , η̃α ) = 0. From (3.10) we obtain the equation for η̃(3) : (3.19) η̃(3) (t) = e−i At η̃0 − e−i At (eiθ z α )(0)e ηα Z t −i A(t−s) isκ[α] d iθ α − e ηα ds (e p )e e ds 0 Z t e−i A(t−s) PcA Fη,3 − eiθ U i −1 η2 η̄ ds + 0 Z t e−i A(t−s) PcA eiθ U i −1 η2 η̄ ds + 0 ≡ η̃1(3) + η̃2(3) + η̃3(3) + η̃4(3) + η̃5(3) . DYNAMICS OF NLS 175 We treat η̃5(3) separately because η2 η̄ is a nonlocal term. Recall that η = U −1 e−iθ η̃. We have the similar decomposition for η, (3.20) η(t) = η(2) (t) + η(3) (t) , η(3) = U −1 e−iθ η̃(3) , η(2) = U −1 e−iθ η̃(2) = U −1 z αe ηα , −1 −iθ (3) η(3) e η̃ j . j =U If we view b and η(2) as order z 2 , and η(3) as order z 3 , we can now decompose F into (3.21) e(3) + F (4) F = F (2) + F (3) + F where F (2) = λQ(2|ζ |2 + ζ 2 ) , F (3) = 2λQ (ζ + ζ̄ )η(2) + ζ η̄(2) + λ|ζ |2 ζ + 2λQ Ra (2) (2ζ + ζ̄ ) , e(3) = 2λQ Rb(2ζ + ζ̄ ) = 2λQ R(a − a (2) )(2ζ + ζ̄ ) , (3.22) F F (4) = 2λQ (ζ + ζ̄ )η(3) + ζ η̄(3) + λQ 2|η|2 + η2 + 2λa Q R(2η + η̄) + 3λa 2 Q R 2 + λ |k|2 k − |ζ |2 ζ . e(3) of terms of order In view of (4.5), F (2) consists of terms of order z 2 ; F (3) and F 3 (4) (3) e z ; and F of higher-order terms. We separate the F term since it depends on b and requires different methods to estimate. From (2.32), we have θ̇ = −[a + (c1 R, Re F)] · [1 + a(c1 R, R) + (c1 R, Re h)]−1 = −[a + (c1 R, Re F)] · [1 − (c1 R, Re h)] + O(z 4 + a 2 + η2 ) (3.23) where = −[a + (c1 R, Re F (2) )] + Fθ,3 + Fθ,4 e(3) + a + (c1 R, Re F (2) ) (c1 R, Re ζ ) Fθ,3 = − c1 R, Re F (3) + F and Fθ,4 = O(z 4 + a 2 + η2 ). 4 Outline of Proofs and Basic Estimates We are now ready to outline the proofs for Theorems 1.1 through 1.4. 4.1 Initial Data and Basic Quantities Under the assumptions of either Theorem 1.1 or Theorem 1.3, we can rewrite the initial data ψ0 using Lemma 2.3 as ψ0 = [Q E0 + h E0 ,0 ]ei2 E0 ,0 , h E0 ,0 ∈ M E0 , for some E 0 , h E0 ,0 , and 2 E0 ,0 , with n = kψ0 kY , kQ E0 k ≥ 9 n 10 , kh E0 ,0 kY ≤ Cε0 n . 176 T.-P. TSAI AND H.-T. YAU We can assume that 2 E0 ,0 = 0 without loss of generality. We now write h E0 ,0 = ζ E0 ,0 + η E0 ,0 according to the spectral decomposition (1.16) in M E0 . Recall the definition of ε in (1.21). We have (4.1) kζ E0 ,0 k ≤ Cεn , kη E0 ,0 kY ≤ Cε2 n 2 , ε ≤ ε0 . For the rest of this paper, we shall take as either given by (1.21) or (1.24). The choice will not be important, and, whenever a specific choice is needed, we shall remark upon it. For all t ≥ 0, we have kψ(t)k L 2 = n. If we write ψ(t) = xφ0 + yφ1 + ξ , ξ ∈ Hc (−1 + V ), then we have |x|2 + |y|2 + kξ k2L 2 = n 2 . Since the spectral projections of −1 + V and L differ by an order of O(λQ 2 ) = O(n 2 ), we have kQ E k = |x| + O(n 3 ), kζ k L 2 = |y| + O(n 3 ), and kηk L 2 = kξ k L 2 + O(n 3 ). In particular, (4.2) kQ E k ≤ 98 n , kζ (t)k L 2 ≤ 98 n , kη(t)k L 2 ≤ 98 n . We now give a list of the expected sizes of frequently used quantities. Let 1 (4.3) 0 ≡ Im 2820 , PcA 820 ≥ λ2 n 2 γ0 , A − 0i − 2κ where 820 = O(n) is defined as in (3.15), and γ0 is the constant from assumption A1. The last statement 0 ≥ λ2 n 2 γ0 will be proven in Lemma 6.1. Recall that {t} ≡ ε−2 n −2 + 0t, {t}−1/2 ∼ min εn, n −1 t −1/2 . (4.4) {t}−1/2 will be the typical size of |z(t)|. The following is a table of order in t for functions: |z(t)| = O(t −1/2 ) , |a(t)| = O(t −1 ) , (4.5) kη(t)kloc = O(t −1 ) , |θ(t)| = O(log t) . The following is a table of order in n for constants: (4.6) Q = O(n) , c1 = O(1) , R = O(n −1 ) , u + = O(1) , u − = O(n 2 ) . The first three estimates in the last equation are due to Lemma 2.1. The last one is because the differences between L + , L − , and H1 = −1 + V − e0 are of order O(λn 2 ), and hence so are the differences between φ1 , u, and v. 4.2 Outlines of the Proofs for Theorems 1.1 and 1.3 We now estimate solutions to the equations (2.30) with the decompositions into main oscillatory and higher-order terms in Section 3. We first fix a time T and let E = E(T ) be the best approximation at the time T so that (1.19) holds at time T , i.e., (4.7) ψ(T, x) = [Q E (x) + h E (T, x)] ei2(T ) . DYNAMICS OF NLS 177 We can now decompose the wave function for all time with respect to this ground state as in (1.18), namely, ψ(t, x) = Q E(T ) (x) + a E(T ) (t)R E(T ) (x) + h E(T ) (t, x) ei2(t) , where h E(T ) (t, x) = ζ E(T ) (t) + η E(T ) (t) ∈ M E(T ) . To simplify the notation, we write a E(T ) = aT , etc. By assumption, we have aT (T ) = 0. We now wish to estimate aT , ζT , and ηT for all time t ≤ T . The following propositions are stated with respect to a decomposition of a fixed nonlinear ground state profile E. We first choose a suitable norm. We need to control the excited-state component z and a local norm of η. We also need a global norm of η to control the nonlocal term η3 . Recall that we can decompose η into a sum of η(2) + η(3) with η(3) decomposed further into a sum of five terms; see (3.20). Since η(2) , η1(3) , and η2(3) are explicit, we need only to control (3) η3−5 (4.8) ≡ 5 X (3) (3) (3) η(3) j = η3 + η4 + η5 , j=3 and we define M(T ) := sup 0≤t≤T (4.9) n {t}1/2 |z(t)| + {t}3/4−σ kη(t)k L 4 + n σ {t} kη(t)k L 2 (3) + εσ (εn)−3/4 {t}9/8 η3−5 (t) L 2 o loc loc where 0 = O(n ) is defined in (4.3) and B22 = + O(n 3 ) will be defined explicitly later in (7.22). For the time being, we need only know that B is a constant with the size given above; hence D = c1 + O(n). If we assume that a(t) is bounded by t −1 up to some time T , we have the following control of M(t). 1 c0 2 1 2 P ROPOSITION 4.1 Suppose z(t) and η(t) are solutions to the equations (2.30) for 0 ≤ t ≤ T . (T can be finite or infinite.) Assume (4.10) |a(t)| ≤ D {t}−1 , 0≤t ≤T. Then we have M(t) ≤ 2 for all t ≤ T . Moreover, if we further assume |z 0 | = εn > 0 and kη0 kY ≤ ε2 n 2 , then |z(t)| ≥ c {t}−1/2 . Recall b(t) := a(t) − a (2) (t). Let ST (t) = b(t) − b(T ) so that t (4.11) a(t) = a(T ) + a20 (z 2 + z̄ 2 ) T + ST (t) . Assuming that a(t) is bounded by t −1 , we can control ST (t) by the following proposition: 178 T.-P. TSAI AND H.-T. YAU P ROPOSITION 4.2 Suppose that M(t) ≤ 2 and |a(t)| ≤ D {t}−1 for 0 ≤ t ≤ T . Recall D = 2B22 / 0 = c1 + O(n). Then we have D −1 D {t} , |a(t)| ≤ |a(T )| + {t}−1 . (4.12) |ST (t)| ≤ 2 2 P ROOF OF T HEOREMS 1.1 AND 1.3: Assuming the previous two propositions, we now prove that (4.13) sup |aT (t)| {t} < D t≤T for all time, where aT is the component with respect to the best approximation at the time T defined in (1.19). From (4.1), the relation (4.13) holds for T = 0. Suppose it holds for time T = T1 . From the continuity of the Schrödinger equation, there exists an δ > 0 (depending on T1 ) such that |E(T )− E(T1 )| ≤ exp[−T1 ](D − supt≤T1 |aT1 (t)| {t}) for all |T − T1 | < δ. Thus the bound (4.13) holds for all |T − T1 | < δ. This proves that the set of such T is open. Now suppose that T1 is the first time that (4.13) is violated. Clearly, we have supt≤T1 |aT1 (t)| {t} ≤ D. From Propositions 4.1 and 4.2, formula (4.13) holds for T1 . Once we have proven (4.13), the conclusions of Propositions 4.1 and 4.2 hold, and they imply both Theorem 1.1 and Theorem 1.3. We remark that the main reason that the previous continuity argument works is due to the explicit identification of the leading terms in a, η, and z. Since these leading terms are essentially computed without using assumptions on the decay of a, the bound on a, i.e., (4.13), is used only to control error terms. We shall prove Theorem 1.4 in Section 7, where we obtain more detailed information about the wave function. The proof of Proposition 4.1 will be given in the next two sections and is followed by a proof of Proposition 4.2. Once again, our strategy of proof is to show that M(0) ≤ 32 and that M(t) ≤ 32 if M(t) ≤ 2. By continuity of M(t), this would imply that M(t) ≤ 32 for all t ≤ T . So for the rest of this and the next sections, we shall freely use that M(t) ≤ 2 and |a(t)| ≤ D {t}−1 . It will follow from Lemmas 5.2 and 5.3 that we can bound {t}3/4−σ kη(t)k L 4 + εσ (εn)−3/4 {t}9/8 η(3) (t) 2 3−5 L loc by 14 . It will follow from Lemma 6.2 that {t}1/2 |z(t)| ≤ (1 + 2σ ) for some small σ . We have thus M(T ) ≤ 32 and conclude Proposition 4.1. We note that, under the assumptions (4.10) and M(T ) ≤ 2, we have, from the equations (2.31) for F, (2.32) for θ̇, (2.30) for ṗ, and (3.9) for b, kFk L 2 ≤ Cn {t}−1 , F − F (2) L 2 ≤ C {t}−3/2 , loc loc (4.14) |θ̇| ≤ C {t}−1 , | ṗ| ≤ Cn {t}−1 , |ḃ| ≤ Cn {t}−3/2 . DYNAMICS OF NLS 179 It follows from (4.14) that |θ(t)| ≤ C log {t} . (4.15) Note that, although θ̇ and ȧ are of the same order, θ̇ is not oscillatory and we cannot expect an estimate better than (4.15). This estimate will not be used in the rigorous proof, but it provides an idea about its size. 5 Estimates of the Dispersive Wave For the dispersive part η, our main interest is its local decay estimate. Due to the presence of the nonlocal term η3 in F, we will also need a global estimate on η. Our goal is to prove, for σ = 1/100, (3) kη(t)k L 4 ≤ 18 {t}−3/4+σ , η3−5 (t) L 2 ≤ Cε−σ (εn)3/4 {t}−9/8 . loc We will need the following calculus lemma: L EMMA 5.1 Let 0 < d < 1 < m, and {t} ≡ ε−2 + 20t. Z t (5.1) |t − s|−d {s}−m ds ≤ Cε2m−2 n 2m+2d−4 {t}−d , 0 Z t (5.2) |t − s|−d {s}−1 ds ≤ Cn 2d−2 {t}−d log(1 + ε2 n 4 t) . 0 If, instead, d ≥ m > 1, Z t (5.3) ht − si−d {s}−m ds ≤ C {t}−m . 0 P ROOF : Denote the first integral by (I). If t ≤ ε −2 n −4 , then {t} ∼ ε−2 n −2 and Z t (I) ∼ |t − s|−d (εn)2m ds . (εn)2m (ε−2 n −4 )1−d 0 = ε2m−2 n 2m+2d−4 (εn)2d ∼ ε2m−2 n 2m+2d−4 {t}−d . If t ≥ ε−2 n −4 , then {t} ∼ 20t and Z t/2 Z −d −m (I) ≤ Ct {s} ds + 0 ≤ Ct t C|t − s|−d C {t}−m ds t/2 −d (ε n ) 0 2 4 m−1 + Ct 1−d {t}−m ≤ Cε2m−2 n 2m+2d−4 {t}−d . Estimate (5.2) is an obvious modification of (5.1). For the last estimate (5.3), denote the second integral by (II). If t ≤ ε−2 n −4 , then {t} ∼ ε−2 n −2 and Z t ht − si−d (εn)2m ds ≤ (εn)2m · C ∼ C {t}−m . (II) ≤ C 0 180 T.-P. TSAI AND H.-T. YAU If t ≥ ε−2 n −4 , then {t} ∼ 0hti and Z t ht − si−d 0 −m hsi−m ds ≤ C0 −m hti−m ∼ C {t}−m . (II) ∼ 0 We conclude the lemma. 5.1 Estimates in L 4 L EMMA 5.2 Suppose that η̃ is given by equation (3.10) and recall η = U −1 e−iθ η̃. Assuming the estimate (4.10) on a(t) and M(T ) ≤ 2, we have kη(t)k L 4 ≤ 1 8 {t}−3/4+σ , where σ = 1/100 if ε is sufficiently small. P ROOF : From the defining equation (3.10) of η̃, we have Z t −3/4 ke η0 kY hti + C|t − s|−3/4 kFη (s)k4/3 ds , η(t)k L 4 ≤ C ke 0 where Fη = eiθ U i −1 [F + θ̇ (a R + ζ )] − eiθ [U, i]θ̇η. By Lemma 2.9, iθ e [U, i]θ̇η(s) 4/3 ≤ C|θ̇ | kη(s)k4 . L Therefore we have kFη (s)k4/3 ≤ C|θ̇| |a| + kζ (s)k4/3 + kη(s)k4 + kF(s)k L 4/3 . From the Hölder inequality and the definition of kF(s)k L 4/3 , we can bound kF(s)k L 4/3 by kF(s)k L 4/3 ≤ C khk2L 4 + |a| khk L 4 + |a|2 + kh 3 k L 4/3 + |a|3 . Since kh 3 k L 4/3 = khk3L 4 and kh(s)k L 4 ≤ kζ (s)k L 4 + kη(s)k L 4 , we have from the assumption MT (t) ≤ 2 that kh(s)k L 4 ≤ C{s}−1/2 + C{s}−3/4 log{s} ≤ C{s}−1/2 . Therefore we have kF(s)k L 4/3 ≤ Cn{s}−1 . Similarly, we have |θ̇(s)| ≤ C{s}−1 . 1 {t}−3/4+σ , we conSince C kη̃0 kY hti−3/4+σ ≤ C kη̃0 kY ε−3/2+2σ {t}−3/4+σ ≤ 16 clude Z t −3/4+σ 1 kη(t)k L 4 ≤ 16 {t} + |t − s|−3/4 Cn{s}−1 ds 0 ≤ 1 16 {t}−3/4+σ + Cnn −1/2 {t}−3/4 log[(εn)2 {t}] ≤ Here we have used Lemma 5.1 for the integration. 1 8 {t}−3/4+σ . DYNAMICS OF NLS 181 5.2 Local Decay Recall that η(3) = U −1 e−iθ η̃(3) and η̃(3) satisfies the equation (3.19). We want to show that η̃(3) is smaller than η̃(2) locally. Recall that the local L 2 norm is defined in (1.7) with r0 the exponent in Lemma 2.7. L EMMA 5.3 Assuming the estimate (4.10) on a(t) and M(T ) ≤ 2, we have (3) η 2 ≤ Cε2 n 2 hti−9/8 , η(3) 2 ≤ C(εn)3/4 {t}−9/8 . (5.4) 1−2 L 3−5 L loc loc −1 Hence we have kηk L 2 ≤ C {t} , and for a local function φ we have loc (φ, |η|2 + |η|3 ) ≤ C hxi2r0 φ ∞ kηk2 2 + kηk 2 kηk2 4 L loc L L L loc (5.5) ≤ C hxi2r0 φ L ∞ {t}−2 . P ROOF : Since U is bounded, it is sufficient to prove the corresponding esti(3) mates for each η̃(3) j , j = 1, 2, . . . , 5, defined in the decomposition (3.19) for η̃ . We first estimate η̃1(3) , (3) η̃ 2 ≤ Ckη0 kY hti−9/8 ≤ Cε2 n 2 hti−9/8 . 1 L loc η̃2(3) η̃3(3) , For and the two terms involving ηα (ηα 6∈ L 2 ) from the definition and the linear estimates in Lemma 2.7, we have (3) η̃ 2 ≤ (εn)2 nhti−9/8 , 2 L loc Z t (3) η̃ 2 ≤ C ht − si−9/8 n 2 {s}−3/2 ds ≤ n 2 (εn)3/4 {t}−9/8 . 3 L loc 0 Here we have used e ηα = O(n), | ṗ| ≤ Cn {t}−1 , and |θ̇| ≤ C {t}−1 in (4.14) and (5.3). Note that, when t is of order 1, η̃1(3) and η̃2(3) are of order η0 and ε2 n 3 , which are larger than C(εn)3/4 {1}−9/8 = O((εn)3 ). Rt We treat η̃4(3) and η̃5(3) together. For η̃4(3) = 0 e−i A(t−s) PcA {Fη,3 −eiθ U i −1 η2 η̄}ds, we can rewrite the integrand using the definition of Fη,3 from (3.13), Fη,3 − eiθ U i −1 η2 η̄ = eiθ U i −1 (F − F (2) − η2 η̄) + θ̇(a R + ζ ) − eiθ [U, i]θ̇η , which consists of only local terms. Since MT (t) ≤ 2, we have Fη,3 − eiθ U i −1 η2 η̄ 4/3 8/7 ≤ C{s}−3/2 . L ∩L Here we have used Lemma 2.9 to estimate eiθ [U, i]θ̇η. Note that, in terms like Qζ η, η is estimated by kηk L 2 ≤ Cn −σ {s}−1 . loc From the Hölder inequality, the L 2 bound (4.2), and the global estimate of η in Lemma 5.2, we have 5/2 1/2 5/2 ≤ C{s}−7/4 , kη3 (s)k L 8/7 ≤ kηk2 kηk4 ≤ Cn 1/2 {s}−3/4+σ 3 kη3 (s)k L 4/3 ≤ kηk34 ≤ C {s}−3/4+σ ≤ C{s}−7/4 . 182 T.-P. TSAI AND H.-T. YAU Rt Thus we have η̃4(3) + η̃5(3) = 0 e−i A(t−s) PcA Fη,3 (s)ds, with Fη,3 (s) 4/3 8/7 ≤ C{s}−3/2 . L ∩L We can bound the L 2loc norm by either the L 4 or L 8 norm. If t > 1, we bound the L 2loc norm by L 8 in 0 ≤ s ≤ t − 1, and by L 4 in t − 1 ≤ s ≤ t. By Lemma 2.7, Z t−1 (3) 1 η̃ + η̃(3) 2 ≤ C kFη,3 (s)k L 8/7 ds 4 5 L loc |t − s|9/8 0 Z t 1 +C kFη,3 (s)k L 4/3 ds |t − s|3/4 t−1 Z t C ht − si−9/8 {s}−3/2 ds + C {t}−3/2 ≤C 0 ≤ C(εn)3/4 {t}−9/8 + C(εn)3/4 {t}−9/8 . In the last line we have used (5.3). If t ≤ 1, we can use the second integral to estimate; hence we get the desired estimate for kη̃4(3) + η̃5(3) k L 2 . loc 6 Excited-State Equation and Normal Form 6.1 Excited-State Equation Let us return to the equation for ṗ from (2.30), (6.1) ṗ = −ieiκt (u + , F) + (u − , F) + (u + , h) + (u − , h) + (u, R)a θ̇ . In view of the decompositions (3.22) of F and (3.23) of θ̇, we can expand the right-hand side of equation (6.1) into terms in order of z: (6.2) ṗ = eiκt cα z α + dβ z β + d1 bz + d2 bz̄ + P (4) . Here α and β are summing over all |α| = 2 and |β| = 3. The coefficients are computed and their properties summarized in the following lemma: L EMMA 6.1 We can rewrite equation (6.1) of ṗ into the form (6.2). The coefficients d1 and d2 and all cα are purely imaginary. P (4) denotes higher-order terms and will be defined in (6.6). Moreover, Re d21 = −0, with 1 A (6.3) 0 = 2820 , Im P 820 ≥ λ2 n 2 γ0 > 0 , A − 0i − 2κ c where γ0 is the constant in assumption A1, and 820 = 5λQφ12 + O(n 3 ) is defined in (3.15). We also have cα = O(n) and dβ , d1 , d2 = O(1). P ROOF : There are two parts in equation (6.1): The part with F and the part with θ̇. We first consider the second part. Recall from (3.23), θ̇ = −a − (c1 R, Re F (2) ) + O(z 3 ) , DYNAMICS OF NLS 183 where O(z 3 ) denotes terms of the order z 3 +az+zη plus higher-order terms. Hence − ieiκt {(u + , h) + (u − , h̄) + (u, R)a}θ̇ = −ieiκt (u + , ζ ) + (u − , ζ̄ ) + O(z 2 ) · − a − (c1 R, Re F (2) ) + O(z 3 ) = −ieiκt (u + , ζ ) + (u − , ζ̄ ) · − a20 (z 2 + z̄ 2 ) − b − (c1 R, Re F (2) ) + O(z 4 ) . The leading terms are of the form z β , bz, and bz̄. These terms are of order O(z 3 ). Since the coefficients in each bracket are real, the coefficients of the leading terms, that is, their contributions to dβ , d1 , and d2 , are all purely imaginary due to the i factor in front. Now we look at the first part of the equation (6.1). The contribution to cα z α is from F (2) = z α φα , −i {(u + , z α φα ) + (u − , z̄ α φα )} . Clearly all coefficients of z α are purely imaginary. Since there is no contribution from the second part to cα z α , we know that the cα are purely imaginary. We also have cα = O(φα ) = O(n). e(3) : The contribution from the first part to d1 bz + d2 bz̄ is from F −i (u + , 2λQ Rb(2ζ + ζ̄ ) + (u − , 2λQ Rb(ζ + 2ζ̄ ) with ζ = zu + + z̄u − . Hence all coefficients of bz and bz̄ are purely imaginary. Together with the analysis of the second part of (6.1), we know d1 and d2 are purely imaginary. The contribution from the first part to dβ z β is from F (3) : −i (u + , F (3) ) + (u − , F̄ (3) ) . Among all dβ , we are only interested in the real part of d21 , which does not come from the second part as we already showed. A coefficient in F (3) has to have an imaginary part in order to have a contribution to Re d21 . The only such source is z 2 η̃20 in η(2) , which lies in the first group of terms in F (3) . Let us call these terms F1(3) : F1(3) = 2λQ (ζ + ζ̄ )η(2) + ζ η̄(2) . Recall that η(2) = U −1 z α η̃α . Denote η0 = z α η̃α and recall (2.24) that U −1 = (U+ )∗ − (U− )∗ C. Hence η(2) = [(U+ )∗ − (U− )∗ C]η0 = (U+ )∗ η0 − (U− )∗ η̄0 , and u + , F1(3) + u − , F̄1(3) = 2λQ, (u + ζ + u ζ̄ )η(2) + 2λQ, (uζ + u − ζ̄ )η̄(2) = 2λQ, (u + ζ + u ζ̄ )[(U+ )∗ η0 − (U− )∗ η̄0 ] + 2λQ, (uζ + u − ζ̄ )[(U+ )∗ η̄0 − (U− )∗ η0 ] 184 T.-P. TSAI AND H.-T. YAU Z = U+ 2λQ(u + ζ + u ζ̄ ) − U− 2λQ(uζ + u − ζ̄ ) η0 d x Z + −U− 2λQ(u + ζ + u ζ̄ ) + U+ 2λQ(uζ + u − ζ̄ ) η̄0 d x . We want to collect terms of the form C z 2 z̄ with Re C 6= 0. However, the only term from η̄0 with a resonance coefficient is z 2 η̃20 , which is of the form z̄ 2 , with two bars. Hence the last integral does not contain z 2 z̄ and is irrelevant. From the first integral, we want to choose z 2 from η0 , i.e., z 2 η̃20 , and choose z̄ from ζ or ζ̄ . The terms with z̄ are U+ 2λQ(u + z̄u − + u z̄u + ) − U− 2λQ(u z̄u − + u − z̄u + ) = 2(U+ φ20 − U− φ02 )z̄ = 2820 z̄ . Here we have used (3.2) and (3.15); hence (6.4) Re d21 = Re(−i) (2820 , η20 ) = Im 2820 , −1 A P 820 = −0 . A − 0i − 2κ c To show 0 ≥ λ2 n 2 γ0 , it suffices to show that (6.5) 820 = 5λQφ12 + O(n 3 ) . Note that 820 = U+ φ20 − U− φ02 = U+ λQφ12 + O(n 3 ) by (3.15) and (3.2). Now U+ = 12 (5A1/2 H −1/2 P1 +5A−1/2 H 1/2 5) is defined in (2.23). Since A1/2 H −1/2 = 1+ O(n 2 ) and A−1/2 H 1/2 = 1+ O(n 2 ) by (9.7) and (9.8), we have U+ = 12 (5P1 + 5) + O(n 2 ). By (2.18), P1 = 5 + O(n 2 ). Therefore we have (6.5). It will follow from the induction assumption that Q = mφ0 + O(n 3 ) with m ≥ 45 n. Then we have, by Lemma 2.8, 1 2 2 2 A 2 0 = 2λ m Im 5φ0 φ1 , P 5φ0 φ1 + O(n 4 ) A − 0i − 2κ c 2 ≥ 2λ2 45 n γ0 + O(n 4 ) . Collecting terms of order O(z 4 + a 2 + η2 ), we have (4) P (4) = −i (u + , F (4) ) + (u − , F ) + [(u + , η) + (u − , η) + (u, R)a]θ̇ − i (u + , ζ ) + (u − , ζ ) (Fθ,3 + Fθ,4 ) . (6.6) 6.2 Normal Form L EMMA 6.2 We can rewrite equation (6.2) of ṗ into a normal form: (6.7) q̇ = δ21 |q|2 q + d1 bq + g , where q is a perturbation of p given in the proof. The coefficient δ21 satisfies (6.8) Re δ21 = Re d21 = −0 . DYNAMICS OF NLS 185 If we assume estimate (4.10) on a(t) and M(T ) ≤ 2, then the error term g(t), to be given by (6.22), satisfies the bound (6.9) |g(t)| ≤ Cε 3/4−σ n 7/4 {t}−13/8 . Furthermore, there is a small positive constant σ such that |q(t)| and |z(t)| are bounded by (6.10) |q(t)| ≤ (1 + σ ) {t}−1/2 , |z(t)| ≤ (1 + 2σ ) {t}−1/2 . If we have |z(0)| = εn + o(εn), we also have lower bound |q(t)| ≥ (1 − σ ) {t}−1/2 , |z(t)| ≥ (1 − 2σ ) {t}−1/2 . P ROOF : From (6.2) we have ṗ = µ cα z α + dβ z β + d1 bz + d2 bz̄ + P (4) , µ = eiκt , and we want to obtain the normal form (6.7). We will repeatedly use the following formula: µm p α d µm p α m α − (6.11) µ p = f α (z) dt iκm iκm where, if α = (α0 α1 ), |α| = α0 + α1 = 2, 3, 4, . . . , (6.12) f α (z) = (α0 + α1 C)( p −1 ṗ) = (α0 + α1 C)z −1 cα z α + dβ z β + d1 bz + d2 bz̄ + P (4) , and C denotes the conjugation operator. The formula is equivalent to integration by parts. We first remove cα z α . Let cα p1 = p − µz α . iκ(1 + [α]) Since [α] is even, 1 + [α] 6= 0. By (6.11) cα ṗ1 = ṗ − cα µz α − µz α f α (z) . iκ(1 + [α]) Decomposing f α (z) into two parts, we can write cα µz α (α0 + α1 C)z −1 cα̃ z α̃ , dβ+ z β = − iκ(1 + [α]) cα g1 = − µz α (α0 + α1 C)z −1 dβ z β + d1 bz + d2 bz̄ + P (4) , iκ(1 + [α]) and we get ṗ1 = δβ µz β + d1 µbz + d2 µbz̄ + µP (4) + g1 with δβ = dβ + dβ+ . Since cα are purely imaginary, hence so are dβ+ , and we have the relation (6.13) Re δβ = Re dβ . 186 T.-P. TSAI AND H.-T. YAU Next we remove d2 µbz̄. Let p2 = p1 − µd2 bz̄ . 2iκ We have ṗ2 = ṗ1 − µd2 bz̄ − µ2 d2 ˙ = µδβ z β + d1 µbz + (µP (4) + g1 + g2 ) , (ḃ p̄ + b p̄) 2iκ where µ2 d2 ˙ . (ḃ p̄ + b p̄) 2iκ Now we deal with δβ z β terms. Let X δβ µz β p3 = p2 − . iκ(1 + [β]) β6 =(21) g2 = − Note 1 + [β] 6= 0 for β 6= (21). We have ṗ3 = ṗ2 − µδβ z β + g3 = δ21 µz 2 z̄ + µd1 bz + (µP (4) + g1 + g2 + g3 ) with g3 = − X δβ µ1+[β] d β (p ) . iκ(1 + [β]) dt β6 =(21) Notice that the equation for p3 has the desired form (6.7), and that the error term (µP (4) + g1 + g2 + g3 ) has the desired decay in t when t is large. However, certain terms in the error term are not small compared to 0z 3 when t is of order 1, and hence need to be treated. (Recall 0 = − Re δ21 .) We will also integrate these terms and include them in the perturbation of p. (3) (3) from P (4) , where η1,2 = η1(3) + η2(3) . Since One such term is of the form Qζ η1,2 the size of η1(3) and η2(3) are comparable to η(2) when the time t is of order 1, and 0z 3 (3) comes from terms of the form Qζ η(2) , we need to treat terms of the form Qζ η1,2 . We will integrate these terms and include them in the perturbation of p. Recall −1 −iθ (3) e e η j , j = 1, 2, and η(3) j =U η0 − (eiθ z α )(0)e e η1(3) + e η2(3) = e−i At e ηα where e η0 = eiθ(0)U η0 , and e ηα , |α| = 2, are defined as in (3.17). Denote χ = (3) (3) (3) e0 = eiθ (0) [U η0 − z α (0)e ηα ] ∈ Hc (A) for the computation η1,2 = η1 + η2 and χ below. We have (6.14) χ (t) = η1(3) + η2(3) (t) = U −1 e−iθ (t) e−i At χ e0 . Recall (6.1). The only source of Qζ χ is from F (4) , where we have a term 2λQ((ζ + ζ̄ )χ + ζ χ). This is the same source for the resonance term Qζ η(2) . Hence terms of the form Qζ χ in µP (4) are exactly Pzη(3) ≡ −iµ u + , 2λQ((ζ + ζ̄ )χ + ζ χ̄) − iµC u − , 2λQ((ζ + ζ̄ )χ + ζ χ̄) . 1,2 DYNAMICS OF NLS 187 Clearly these terms can be summed in the form Pzη(3) = µ(zφ + z̄φ, χ ) + Cµ−1 (zφ + z̄φ, χ ) . (6.15) 1,2 Here each φ = O(n) stands for a different local smooth function. Recall U −1 = U++ − U−∗ C from (2.24); hence U −1 (z f + z̄g) = z(U+∗ f − U−∗ ḡ) + z̄(U+∗ g − U−∗ f¯) (cf. (7.14)). Together with (6.14) we have Z pφ1 + µ2 p̄φ2 e−iθ e−i At χ e0 d x Pzη(3) = 1,2 Z (6.16) e0 d x +C µ−2 pφ3 + p̄φ4 e−iθ e−i At χ for some local functions φ1 , φ2 , φ3 , φ4 = O(n). Recall µ(s) = eiκs . Define Z tZ f 1 (t) = φ1 e−i As χ e0 d x ds , ∞ Z tZ f 2 (t) = φ2 ei2κs e−i(A−0i)s χ e0 d x ds , ∞ Z tZ φ3 e−i2κs e−i As χ f 3 (t) = e0 d x ds , ∞ Z tZ f 4 (t) = e0 d x ds . φ4 e−i As χ ∞ η20 = These functions in t depend on the initial data only. Recall χ e0 contains e −1 A −(A − 0i − 2κ) Pc 820 , which is not a local term. It is clear by Lemma 2.7 that Z t | f j (t)| ≤ Cnhsi−9/8 ke χ0 k ds ≤ Cε 3/2 n 2 hti−1/8 . ∞ However, we claim we have | f j (t)| ≤ Cε3/2 n 2 hti−9/8 , (6.17) j = 1, 2, 3, 4 . In fact, (6.17) is clear for f j (t), j 6= 2, since we can integrate them explicitly. For example, Z f 1 (t) = (−i A)−1 (PcA φ1 )e−i At χ e0 d x = O(hti−9/8 ) , by Lemma 2.7. Here we have used that A−1 is bounded in L r2 ∩ Hc (A). The e0 is problem for the term f 2 (t) is that the factor ei2κs e−i As gives resonance while χ not a local term. However, the main term in f 2 that concerns us is Z tZ Z i2κs −i(A−0i)s φ2 e e e η20 d x ds = −i φ2 (A − 2κ − 0i)−2 PcA 820 d x , ∞ which is still of order O(hti−9/8 ) by Lemma 2.7. Thus we also have (6.17) for f 2 . 188 T.-P. TSAI AND H.-T. YAU We can now integrate Pzη(3) : 1,2 Pzη(3) = (6.18) 1,2 d ( p4 ) + g4 , dt where p4 is of similar form to Pzη(3) , 1,2 f 2 + C pe−iθ f 3 + p̄e−iθ f 4 , (6.19) d d d d −iθ −iθ −iθ −iθ ( pe ) f 3 + ( p̄e ) f 4 . g4 = ( pe ) f 1 + ( p̄e ) f 2 + C dt dt dt dt By (6.17), p4 = pe (6.20) −iθ f 1 + p̄e −iθ | p4 | ≤ Cε3/2 n 2 hti−9/8 |z| , |g4 | ≤ Cε3/2 n 2 hti−9/8 (|z θ̇| + | ṗ|) . (3) . We have integrated terms of the order Qzη1−2 Other terms of concern for t = O(1) are related to a R and η from P (4) . (We are not worried about terms from g1 , g2 , and g3 , since the coefficients of these remainder terms of integration by parts are multiplied by O(n); see (4.14).) These terms are of the form Pa 2 ,aη = µ (φ, z 2 a R) + (φ, Q(a R)2 ) + (φ, Qa Rη) + (φ, za Rη) , where φ denotes some local functions. Since a = O({t}−1 ), R = O(n −1 ), and η = O({t}−1 ) locally, the largest terms here are of order n −1 {t}−2 , which is larger than 0z 3 when t is small. However, terms linear in η can be integrated as A(zη) in the next section. The first two terms, µz 2 a R and µQ(a R)2 , are multiplied by µ = eiκt ; hence they are oscillatory and can be integrated as cα z α terms and as A(zb) in the next section. (Note a = a20 (z 2 + z̄ 2 ) + b and all terms of the form µz γ bm with |γ | + 2m = 4 are oscillatory; i.e., they have a nonzero phase.) Since the computation is the same as for A(zη) , cα z α , and A(zb) , we shall skip the computation and just give the result: d Pa 2 ,aη = ( p5 ) + g5 , dt where p5 is of similar form to Pa 2 ,aη and | p5 | ≤ Cn −1 {t}−2 , (6.21) |g5 | ≤ C {t}−5/2 . Now we let q = p3 − p4 − p5 . We have q̇ = δ21 µz 2 z̄ + µd1 bz + µP (4) + g1 + g2 + g3 − Pzη(3) + g4 − Pa 2 ,aη + g5 1,2 2 (4) = δ21 | p| p + d1 bp + µP − Pzη(3) − Pa 2 ,aη + g1 + g2 + g3 + g4 + g5 1,2 = δ21 |q|2 q + d1 bq + g , where (6.22) g= µP (4) − Pzη(3) − Pa 2 ,aη + g1 + · · · + g5 1,2 + δ21 (| p| p − |q|2 q) + d1 b( p − q) . 2 DYNAMICS OF NLS 189 Hence we have arrived at the normal form. (Note that, although we kept ṗ and ḃ in the definition of g2 , it should be replaced by their corresponding equations (6.2) and (7.2)). Also, note we have X cα δβ µz β µd2 bz̄ (6.23) q = p− µz α − − − p4 . iκ(1 + [α]) 2iκ iκ(1 + [β]) β6 =(21) Finally, from the explicit form of g(t), it is bounded by terms of the form (3) nz 4 + nb2 + (nφ, η2 ) + nzη3−5 . Assuming the bounds on a(t) and M(t), we get the bound (6.9) for g(t): |g(t)| ≤ Cε 3/4−σ n 7/4 {t}−13/8 . To conclude this lemma, it remains to prove the estimate on q, which follows from the next lemma. Our main restriction on the size of η0 comes from the term (nφ, (η1(3) )2 ) in g(t). Since we want it to be smaller than 0|z|3 , we need η0 ε3/2 n 2 for t = O(1). We assume η0 ≤ ε2 n 2 for simplicity. 6.3 Decay Estimates In this subsection we present a calculus lemma that deals with the decay of q(t). We will write (6.24) q(t) = ρ(t)eiω(t) , where ρ = |q| and ω is the phase of q. Recall {t} = ε−2 n −2 + 20t , {t} ∼ max{ε −2 n −2 , t} . Before we proceed with the proof, we give some simple facts of an ordinary differential equation. Example. Consider real functions r (t) > 0 that solve (6.25) ṙ (t) = −r (t)3 − ε(1 + t)−3 for t > 0, where ε > 0 is small. We have the following facts: (1) All solutions r (t) satisfy r (t) ≤ (C + 2t)−1/2 with C = r (0)−2 . (2) There is a number r1 > 0 such that if r (0) > r1 , then r (t) ∼ (C + 2t)−1/2 . (3) There is a unique global solution r0 (t) of (6.25) such that r0 (t) ∼ t −2 as t → ∞ . (4) If r (0) < r0 (0), then r (t) = 0 in finite time. (5) If r (0) > r0 (0), then Z ∞ r (s)2 ds = ∞ . 0 190 T.-P. TSAI AND H.-T. YAU Thus the behavior of r (t) depends on which term on the right side of (6.25) dominates. L EMMA 6.3 Recall that n 0 , ε0 , n, and ε given in Theorems 1.1 through 1.4 satisfy 0 < n ≤ n 0 and 0 < ε ≤ ε0 . Let 0 ≈ n 2 and σ = 0.01. Suppose a positive function ρ(t) satisfies ρ̇ = −0ρ 3 + e g (t) , (6.26) where the error term |e g (t)| ≤ Cε 3/4−σ n 7/4 {t}−13/8 , {t} = {t}ε = ε−2 n −2 + 20t . (i) Suppose ρ0 = εn. Then there is a constant m = m(ε0 ) > 1 such that m −1 {t}−1/2 ≤ ρ(t) ≤ m {t}−1/2 . Moreover, m(ε0 ) → 1+ as ε0 → 0+ . (ii) Suppose that ρ(0) ≤ εn. Then we have ρ(t) ≤ m {t}−1/2 . The above example shows that we cannot expect a lower bound for ρ(t) under the assumption of part (ii), when ρ(0) is too small compared with the error term. P ROOF : We first prove part (i). Let ρ+ = m {t}−1/2 and ρ− = m −1 {t}−1/2 , with m > 1 to be determined. We have ρ+ (0) > ρ(0) > ρ− (0). Moreover, ρ̇+ = −0m −2 ρ+3 = −0ρ+3 + 0(1 − m −2 )ρ+3 ≥ −0ρ+3 + e g if 0(1 − m −2 )m 3 {t}−3/2 ≥ Cε3/4−σ n 7/4 {t}−13/8 . Also, ρ̇− = −0m 2 ρ−3 = −0ρ−3 − 0(m 2 − 1)ρ−3 ≤ −0ρ−3 + e g if 0(m 2 − 1)m −3 {t}−3/2 ≥ Cε3/4−σ n 7/4 {t}−13/8 . Since {t}−3/2 ≥ (εn)1/4 {t}−13/8 , both inequalities hold if 0(1 − m −2 )m 3 , 0(m 2 − 1)m −3 ≥ Cε1−σ n 2 . Since 0 ≈ n 2 and ε ≤ ε0 , the above is true for m > 1 arbitrarily close to 1 by choosing ε0 sufficiently small. By comparison, we have ρ− (t) < ρ(t) < ρ+ (t) for all t. The upper bound in part (ii) follows from the same proof by comparing ρ(t) with ρ+ (t). DYNAMICS OF NLS 191 7 Change of the Mass of the Ground State Recall that a(t) satisfies the integral equation (3.3), and we have derived the main oscillatory terms of a(t) in (3.8): a(t) = a (2) (t)+b(t) and a (2) = a20 (z 2 + z̄ 2 ), with b(t) given by (3.9); that is, Z t a(t) = a(T ) + (7.1) (c1 Q, Im(F + θ̇ h))ds T Z t (2) (2) = a(T ) + a (t) − a (T ) + ḃ(s)ds , T where c1 = (Q, R) (7.2) −1 and ḃ = c1 Q, Im[F − F (2) + θ̇ h] − 4a20 Re e−2iκs p ṗ . (2) (3) 2 2 Note that, after P the3substitution η = η + 4η and2 a = a20 (z + z̄ ) + b, ḃ is of the form Re {C z + C zb + (zφ, η) + C z + C z b} plus error. We will perform several integrations by parts to arrive at the form Z t −1 a(t) = O(t ) + O(t −2 )ds = O(t −1 ) T −1 and obtain the estimate |a(t)| ≤ O(t ). We first decompose the integrand of a(t) in (7.1) according to order in z; see table (4.5). Recall η = η(2) + η(3) and a = a20 (z 2 + z̄ 2 ) + b. Also, recall from (3.23) that (7.3) θ̇ = − a (2) + b + (c1 R, Re F (2) ) + Fθ,3 + Fθ,4 e(3) + a (2) + b + (c1 R, Re F (2) ) (c1 R, Re ζ ) , Fθ,3 = − c1 R, Re F (3) + F and Fθ,4 = O(z 4 + a 2 + η2 ). Thus we have (c1 Q, Im θ̇ h) = θ̇ c1 Q, Im ζ + η(2) + η(3) (7.4) = − a (2) + b + (c1 R, Re F (2) ) (c1 Q, v) Im z + Fθ,3 (c1 Q, v) Im z − a (2) + b + (c1 R, Re F (2) ) c1 Q, Im η(2) (7.5) + Fθ,4 (c1 Q, v) Im z + Fθ,3 + Fθ,4 c1 Q, Im η(3) (7.6) + θ̇ c1 Q, Im η(3) . Here the second line (7.4) is of order O(z 2 ) and the third line (7.5) of order O(z 3 ), and the last line (7.6) contains higher-order terms. Together with the decompositions (3.22) of F we can decompose the integrand of a(t) in (7.1) according to order in z, (c1 Q, Im(F + θ̇ h)) = A(2) + A(3) + A(4) + A(5) , A(3) = [A(zb) + A(zη) + A(z ) ] , 3 192 T.-P. TSAI AND H.-T. YAU where A(2) consists of O(z 2 ) terms, A(3) of O(z 3 ) terms, A(4) of O(z 4 ) terms, and A(5) of higher-order terms. They are given explicitly by A(2) = (c1 Q, Im λQζ 2 ) , A(zb) = (c1 Q, Im 2λQ Rbζ ) − b(c1 Q, v) Im z , A(zη) = (c1 Q, Im 2λQζ η) , 3 A(z ) = c1 Q, Im λ|ζ |2 ζ + 2λQ Ra (2) ζ − a (2) + (c1 R, Re F (2) ) (c1 Q, v) Im z , A(4) + A(5) = c1 Q, Im λQη2 + 2λQ Raη + λ |a R + ζ |2 (a R + ζ ) − |ζ |2 ζ + (7.5) + (7.6) . (7.7) A (4) A(5) = (7.5) + c1 Q, Im λQ(η(2) )2 + 2λQ Raη(2) + λ 2|ζ |2 (a R + η(2) ) + ζ 2 (a R + η(2) ) , = c1 Q, Im λQ[2η(2) η(3) + (η(3) )2 ] + 2λQ Raη(3) + c1 Q, Im λ 2|ζ |2 η(3) + ζ 2 η(3) + `2 ζ̄ + 2|`|2 ζ + `2 `¯ + (7.6) , ` = aR + η . Since ζ = zu + + z̄u − , we have Im ζ = Im z(u + − u − ) and Im ζ 2 = (Im z 2 )(u 2+ − u 2− ) . We now proceed to integrate them term by term. We have already integrated A(2) in Section 3.1. Write (7.8) e−iκt ṗ = cα z α + dβ z β + d1 bz + d2 bz̄ + P (4) := cα z α + P (3,4) , where P (3,4) is defined by the last equality. We can rewrite our result in Section 3.1 as Z t Z t (2) (2) (2) (7.9) A ds = a (t) − a (T ) + A2,3 + A2,4 + A2,5 ds , T T a (2) = a20 (z 2 + z̄ 2 ) , a20 = (4κ)−1 c1 Q, λQ(u 2+ − u 2− ) = O(n 2 ) , (7.10) A2,3 = −4a20 Re zcα z α , A2,4 = −4a20 Re dβ zz β + d1 bz 2 , A2,5 = −4a20 Re z P (4) . Note here we have used Re zd2 bz̄ = 0 from Lemma 6.1. We can also rewrite (7.2), the equation for b = a − a (2) , as (7.11) 3 ḃ = A(zb) + A(zη) + A(z ) + A2,3 + A(4) + A2,4 + A(5) + A2,5 . DYNAMICS OF NLS 193 7.1 Integration of O(z 3 ) Terms There are three terms of order O(z 3 ): A(zb) , A(zη) , and A(z ) (we will absorb 3 A2,3 into A(z ) ). We first integrate A(zb) . From its explicit form we have A(zb) = C2 b Im z for a real constant C2 = O(n). Hence, using the integration-by-parts formula (6.11) and the decompositions (7.8) and (7.11), Z t Z t Z t 1 (zb) −iκs d A ds = C2 b Im z ds = C2 Im e bz − (bp) ds −iκ dt T T T Z t C2 = z ḃ + b(cα z α + P (3,4) )ds 2 Re zb − 2κ T Z t A zb,4 + A zb,5 ds , = czb b(z + z̄) + 3 T where P (3,4) is defined in (7.8) and C2 = O(n) , 2κ 3 = −czb (z + z̄) A(zb) + A(zη) + A(z ) + A2,3 − 2czb b Re cα z α , czb = (7.12) A zb,4 A zb,5 = −czb (z + z̄) A(4) + A2,4 + A(5) + A2,5 − 2czb b Re P (3,4) . We now integrate A(zη) . Recall from (2.24) that U = U+ + U− C and U −1 = U+∗ − U−∗ C. We will also use the following formulae: Z Z Re d x f (Cg) = Re d x(C f )g , Z Z (7.13) Im d x f (Cg) = − Im d x(C f )g , (7.14) U (z f + z̄g) = z(U+ f + U− ḡ) + z̄(U+ g + U− f¯) . Recall η = U −1 e−iθ η̃. Denote η0 = e−iθ η̃ and hence η = U −1 η0 . We have from (7.13) and (7.14) the identity A(zη) = (c1 Q, Im 2λQζ η) Z = Im d x2c1 λQ 2 ζ η Z = Im d x2c1 λQ 2 (zu + + z̄u − )(U+∗ − U−∗ C)η0 Z = Im d x (U+ + U− C)[z(2c1 λQ 2 u + ) + z̄(2c1 λQ 2 u − )] η0 Z = Im d x(zφ3 + z̄φ4 )η0 , 194 T.-P. TSAI AND H.-T. YAU where φ3 = U+ (2c1 λQ 2 u + ) + U− (2c1 λQ 2 u − ) , φ4 = U+ (2c1 λQ 2 u − ) + U− (2c1 λQ 2 u + ) . We now rewrite (7.15) η0 (s) = e−iθ η̃ = e−iθ e−i As f (s) , Z s i As f (s) = e η̃ = η̃0 + eiτ A PcA Fη (τ )dτ . 0 The reason we work with f instead of η̃ is that those terms of the same order (z 2 ) in ∂s f (s) = eis A PcA Fη (s) are explicit. We now have Z t Z t A(zη) ds = Im (φ3 , zη0 ) + (φ4 , z̄η0 )ds T T Z t φ3 , e−is(A+κ) pe−iθ f ds = Im T Z t + Im φ4 , e−is(A−κ) p̄e−iθ f ds . T Note f (s) ∈ Hc (A). We first compute the first integral, which is equal to t 1 −is(A+κ) −iθ pe f = Im φ3 , e −i(A + κ) T Z t 1 d −is(A+κ) −iθ − Im e pe f ds φ3 , −i(A + κ) ds T t 1 A −iθ = Re P 5φ3 , ze η̃ A+κ c T Z t 1 A −iκs 0 0 −iθ A − Re ṗη − i θ̇ zη + ze Pc Fη ds P 5φ3 , e A+κ c T Z t t φ5 , e−iκs ṗη0 − i θ̇ zη0 + ze−iθ PcA Fη ds = Re φ5 , zη0 T − Re T where 1 P A 5φ3 . A+κ c We are careful in adding PcA 5 so that φ5 makes sense. We can do so since f (s) ∈ Hc (A). Similarly, the second integral is equal to Z t t = Re φ6 , z̄η0 T − Re φ6 , eiκs ṗη0 − i θ̇ z̄η0 + z̄e−iθ PcA Fη ds φ5 = T with φ6 = 1 P A 5φ4 . A−κ c DYNAMICS OF NLS 195 Rt Using η0 = U η, we can rewrite the leading terms of T A(zη) ds in the form Z Re(φ5 , zη0 ) + Re(φ6 , z̄η0 ) = Re d x(zφ5 + z̄φ6 )(U+ + U− C)η Z = Re d x[(U+∗ + U−∗ C)(zφ5 + z̄φ6 )]η Z = Re d x(zφ8 + z̄φ7 )η = Re(zφ7 + z̄φ8 , η) , where we have used (7.13) and (7.14) again, with the convention (2.1) and φ8 = U+∗ φ5 + U−∗ φ6 , φ7 = U+∗ φ6 + U−∗ φ5 . Tracking our definition, we have φi = O(n 2 ), i = 3, 4, . . . , 8. The remaining integral has the integrand = − Re φ5 , e−iκs ṗη0 − i θ̇η0 + ze−iθ PcA Fη − Re φ6 , eiκs ṗη0 − i θ̇ z̄η0 + z̄e−iθ PcA Fη Z = − Re d x e−iκs ṗφ5 + eiκs ṗφ6 U η + (zφ5 + z̄φ6 ) −i θ̇U η + e−iθ PcA Fη . Using (3.11) we have −i θ̇U η + e−iθ PcA Fη = PcA U i −1 [F + θ̇(a R + ζ + η)] . Thus the integrand of the remaining integral can be written as A zη,3 + A zη,4 + A zη,5 , and we have Z t Z t t (zη) (7.16) A ds = [Re(zφ7 + z̄φ8 , η)]T + A zη,3 + A zη,4 + A zη,5 ds , T Z A zη,3 = − Re Z T d x(zφ5 + z̄φ6 )PcA U i −1 z α φα d x cα z α φ5 + cα z α φ6 U η(2) e(3) + (zφ5 + z̄φ6 )PcA U i −1 F (3) + F (2) − a + b + c1 R, Re F (2) ζ Z = − Re d x cα z α φ5 + cα z α φ6 U η(3) + P (3,4) φ5 + P (3,4) φ6 U η A zη,4 = − Re (7.17) A zη,5 + (zφ5 + z̄φ6 )PcA U i −1 F (4) + θ̇(a R + η) + (Fθ,3 + Fθ,4 )ζ , where P (3,4) stands for the higher-order terms in e−iαs ṗ (7.8), and we have used the definitions of F (3.22) and θ (7.3). We observe that the only appearance of θ 196 T.-P. TSAI AND H.-T. YAU is as the exponent of eiθ . Also, A zη,3 is a sum of monomials of the form cz β with |β| = 3. There is no a or η in A zη,3 . Summarizing our effort, we have obtained t a(t) = a(T ) + a20 z 2 + z̄ 2 + czb b(z + z̄) + Re(zφ7 + z̄φ8 , η) T Z t (z 3 ) (7.18) A + A2,3 + A(zη,3) + A(4) + A2,4 + A(zb,4) + A(zη,4) + T + A(5) + A2,5 + A(zb,5) + A(zη,5) ds . All terms in [A(z ) + A2,3 + A(zη,3) ] are of the form cz β , |β| = 3. We define Aβ so that 3 Aβ z β = A(z ) + A2,3 + A(zη,3) 3 = c1 Q, Im λ|ζ |2 ζ + 2λQ Ra (2) ζ − a (2) + c1 R, Re F (2) (c1 Q, v) Im z Z − 4a20 Re zcα z α − Re d x(zφ5 + z̄φ6 )PcA U i −1 z α φα . From this explicit form, we have Aβ = O(n). By integration by parts, we have Z t Z t Aβ Aβ z β ds = aβ z β − aβ z β f β (z) ds , aβ := = O(n) , i[β]κ T T Z t β = aβ z + (7.19) A3,4 + A3,5 ds T where f β is defined in (6.12) and X A3,4 = (7.20) −aβ z β (β0 + β1 C)z −1 cα z α β=(β0 ,β1 ),β0 +β1 =3 A3,5 = X −aβ z β (β0 + β1 C)z −1 P (3,4) . β=(β0 ,β1 ),β0 +β1 =3 Notice that [β] 6= 0 and aβ = aβ̄ . Substituting (7.19) into (7.18), we obtain t a(t) = a(T ) + a20 z 2 + z̄ 2 + czb b(z + z̄) + Re(zφ7 + z̄φ8 , η) + aβ z β T Z t (4) + (7.21) A + A2,4 + A(zb,4) + A(zη,4) + A3,4 T + A(5) + A2,5 + A(zb,5) + A(zη,5) + A3,5 ds . We have finished the integration of all terms of order O(z 3 ) in equation (7.1) for a(t). Thus there are no order-O(z 3 ) terms in the integrand in (7.21). The order z 4 terms are collected in first group of integrands in (7.21), and the higher-order terms in the second group. DYNAMICS OF NLS 197 In principle, order-z 4 terms have the following forms: z 4 , b2 , η2 , z 2 b, z 2 η, bη, and zη(3) . Closer examination shows that b2 , b|z|2 , or zη(3) never occur. (Recall that zη(3) is part of zη, which we already treated when we integrated A(zη) .) Furthermore, the terms involving η can be removed by making the substitution η = η(2) + η(3) . A precise statement on these integrands is given in the following lemma. It is crucial that no term involving b2 appear inR the integrand. Otherwise, t it will result in an inequality of the form b(t) ≤ C/t + b(s)2 ds, which does not guarantee b(t) ≤ C/t. L EMMA 7.1 The integrands of order O(z 4 ) in (7.21) can be summed into the form A(4) + A2,4 + A(zb,4) + A(zη,4) + A3,4 = B22 |z|4 + Re A40 z 4 + A31 z 3 z̄ + Ab2 bz 2 . There are no terms of the form b2 , b|z|2 , or zη(3) . Moreover, we have (7.22) B22 = 12 c1 0 + O(n 3 ) , A40 , A31 = O(n) , Ab2 = O(n) . P ROOF : Recall the definitions (7.7), (7.10), (7.12), (7.17), and (7.20): A(4) = c1 Q, Im λQ(η(2) )2 + 2λQ Raη(2) + λ 2|ζ |2 (a R + η(2) ) + ζ 2 (a R + η(2) ) + Fθ,3 (c1 Q, v) Im z − a (2) + b + c1 R, Re F (2) (c1 Q, Im η(2) ) A2,4 = −4a20 Re dβ zz β + d1 bz 2 3 A(zb,4) = −czb (z + z̄) A(zb) + A(zη) + A(z ) + A2,3 − 2czb b Re cα z α Z A(zη,4) = − Re d x cα z α φ5 + cα z α φ6 U η(2) e(3) − a (2) + b + c1 R, Re F (2) ζ + (zφ5 + z̄φ6 )PcA U i −1 F (3) + F X A3,4 = −aβ z β (β0 + β1 C)z −1 cα z α . β=(β0 ,β1 ),β0 +β1 =3 The first part of the lemma is obtained by direct inspection. Note we can deal with A zη,4 in the same manner as we dealt with A(zη) . The orders of the coefficients are also obtained by direct check with the following table in mind: Q = n, R = n −1 , a20 = n , 2 c1 = 1 , aβ = n , ζ ∼ z + n 2 z̄ , cα = n , czb = n , a R ∼ n −1 ρ 2 , dβ , d1 , d2 = 1 , φ7 , φ 8 = n 2 η ∼ nρ 2 . 198 T.-P. TSAI AND H.-T. YAU We note that all C|z|4 terms with C = O(n 2 ) but one are killed by the Im operator. The only surviving term is the following resonance term of A(4) : c1 Q, Im λζ 2 η(2) = X O(n 2 )z γ . c1 Q, Im λz 2 u 2+ z 2 η20 + O(n 3 )|z|4 + |γ |=4,γ 6 =(22) The first term on the right side is equal to 12 c1 0|z|4 + O(n 3 ), and this is the main term in B22 . This proves Lemma 7.1. 7.2 Integration of Higher-Order Terms Next we proceed to integrate out those oscillatory terms in the integrand of (7.21) for a(t). The first group consists of terms of the form z γ , |γ | = 4. Z t Re A40 z 4 + A31 z 3 z̄ + Ab2 bz 2 ds T A40 z 4 A31 z 3 z̄ = Re + + −4iκ −2iκ Z t A40 z 4 − Re f 40 (z) + T −4iκ where d (bp 2 ) ds t Ab2 bz 2 −2iκ T A31 z 3 z̄ Ab2 µ−2 d f 31 (z) + (bp 2 )ds −2iκ −2iκ ds = ḃ p 2 + 2bp ṗ = O(z 5 ). Let A40 A31 Ab2 = O(n) , a31 = = O(n) , ab2 = = O(n) , −4iκ −2iκ −2iκ d A4,5 = − Re a40 z 4 f 40 (z) + a31 z 3 z̄ f 31 (z) + ab2 (bp 2 ) . ds Then we have the identity Z t Re A40 z 4 + A31 z 3 z̄ + Ab2 bz 2 ds = T Z t 4 3 2 t Re a40 z + a31 z z̄ + ab2 bz T + A4,5 ds . a40 = T After this integration, the integrand in (7.21) becomes A(5) + A2,5 + A(za,5) + A(zη,5) + A3,4 + A4,5 . Since |A4,5 | is of order z 5 , the integrands are of order z 5 or higher. More precisely, they are of the orders (3) (3) + z 2 η3−5 + ··· . n z 5 + z(a R)2 + z(a R)η + (a R)2 η + zη2 + z 2 η1−2 (The first n is from the c1 Q in equation (2.30) for ȧ.) These terms are of order t −2−1/8 and thus can be considered as error terms when t large. Unfortunately, we need to control their behavior even for t of order 1. In this region, their orders in terms of n are crucial. DYNAMICS OF NLS 199 We first note that nz 5 is bounded by εn 2 |z|4 |B22 ||z|4 . Since kηk L 2 −1 C {t} , the term nzη is also much smaller than |B22 ||z| . The term 2 4 loc (3) nz 2 η3−5 ≤ is −2−1/8 |B22 ||z|4 . Thus the main trouble lies in terms bounded by ε−σ n(εn)3/4 {t} of order (3) nz(a R)2 + nz(a R)η + n(a R)2 η + nz 2 η1−2 . These terms have faster time-decay order than |B22 ||z|4 , namely, t −2−1/8 , but they are larger than |B22 ||z|4 when time t = O(1). They are oscillatory terms of order z 5 that we will integrate. (In contrast, terms of order z 2k may be nonoscillatory.) Since the integration procedures for these terms are the same as those for O(z 3 ) terms, we only sketch the main steps: (1) (2) (3) (4) (5) Replace all a in the above group by a20 (z 2 + z̄ 2 ) + b. 3 All terms z γ with |γ | = 5 are oscillatory and can be integrated as A(z ) . nz 3 b R 2 and nzb2 R 2 are also oscillatory and can be integrated as A(zb) . All terms linear in η—z 3 η, zbη, and b2 η—can be integrated as A(zη) . (3) can be integrated as Pzη(3) , defined in (6.15). Specifically, we have z 2 η1−2 1−2 terms of the form X d (3) A z 2 η(3) = Re nφ , z α η1−2 az 2 η(3) + A z 2 η(3) ,5 = 1−2 1−2 1−2 dt |α|=2 where φ = O(1) are some complex local functions and az 2 η(3) and A z 2 η(3) ,5 1−2 1−2 are defined similarly to p4 and g4 in (6.19) with the decay a 2 (3) ≤ C {t}−1 ε2 n 3 hti−9/8 , z η1−2 A 2 (3) ≤ C {t}−3/2 ε2 n 3 hti−9/8 ≤ Cε|B22 | {t}−2 . z η ,5 1−2 Note that, as in p4 and g4 , we have terms of the form (φ, (A − 2κ − 0i)−2 PcA 8)nz 2 in az 2 η(3) , which still has the claimed decay by Lemma 2.7. 1−2 To summarize, we have obtained t a(t) = a(T ) + a20 z 2 + z̄ 2 + czb b(z + z̄) + Re(zφ7 + z̄φ8 , η) + aβ z β T t + Re a40 z 4 + a31 z 3 z̄ + ab2 bz 2 + az 2 η0 T (7.23) (3) t + Re C z 5 + C z 3 b + C zb2 + C z 3 η + C zbη + Cb2 η + C z 2 η1−2 T Z t + B22 |z|4 + B5 ds , T 1 c0 2 1 + O(n 3 ), the terms in the third line denote various terms of where B22 = similar form (e.g., C z 5 means C z γ with |γ | = 5), and (7.24) |B5 (t)| ≤ C(ε + n)σ n 7/4 {t}−2−1/8 0 {t}−2 , 200 T.-P. TSAI AND H.-T. YAU assuming the bound (4.10) on a(t) and M(T ) ≤ 2. Recall b(t) := a(t) − a (2) (t) and ST (4.11) is defined by t a(t) = a(T ) + a20 z 2 + z̄ 2 T + ST (a, z, η, θ )(t) . From (7.23), ST can be viewed as a function of a, z, η, and θ and explicitly given by ST (a, z, η, θ)(t) t := czb b(z + z̄) + Re(zφ7 + z̄φ8 , η) + aβ z β T t + Re a40 z 4 + a31 z 3 z̄ + ab2 bz 2 + az 2 η0 T (7.25) (3) t + Re C z 5 + C z 3 b + C zb2 + C z 3 η + C zbη + Cb2 η + C z 2 η1−2 T Z t + B22 |z|4 + B5 ds . T The right-hand side depends on b, η(3) , etc., which are not explicitly given as variables in ST . But all these variables can be traced back to the basic variables a, z, η, and θ, e.g., b = a − a20 (z 2 + z̄ 2 ). Making the same assumptions as in Proposition 4.1, we have B22 −1 D {t} ≤ o(1) + {t}−1 . (7.26) |ST (a, z, η, θ)(t)| ≤ C1 (D) {t}−3/2 + 20 4 We also have a(t) = a(T )+[a20 (z 2 + z̄ 2 )]tT + S(a)(t) = a(T )+[o(1)+ D/4] {t}−1 . We conclude with the following lemma: L EMMA 7.2 Suppose that M(t) ≤ 2 and |a(t)| ≤ D {t}−1 for 0 ≤ t ≤ T . Recall D = 2B22 / 0 = c1 + O(n). Then we have (7.27) |ST (a, z, η, θ)(t)| ≤ D −1 {t} , 2 |a(t)| ≤ |a(T )| + D −1 {t} . 2 This lemma gives Proposition 4.2 and concludes the proof of Theorems 1.1 and 1.3. Since |E 0 − E j+1 | ≤ |E 0 − E j |+|γ | ≤ 98 Dε2 n 2 +|γ | ≤ 2Dε2 n 2 , by Lemma 2.4 again with E 1 standing for E 0 and E 2 standing for E j+1 , (7.28) |E 0 − E j+1 − a j+1 (0)| ≤ Cn −1 (εn)(2Dε2 n 2 ) = Cε3 n 2 . The constant C here is independent of j since we use a bound for |E 0 − E j+1 | that is independent of j. Since |a j+1 (0)| ≤ 12 Dε2 n 2 , we have (7.29) |E 0 − E j+1 | ≤ Dε2 n 2 . From this, we also get kψ0 − Q j+1 ei2j+1 (0) kY ≤ 2εn. We have thus proven the induction hypothesis (I j+1 ) for t = Tj+1 . The induction proof is complete. DYNAMICS OF NLS 201 P ROOF OF T HEOREM 1.4: Suppose the assumption of Theorem 1.4 holds. The existence of E ∞ and the upper bound (1.22) follows from the same proof for Theorem 1.3. Recall from (1.24) that |z E0 ,0 | = εn , kη E0 ,0 kY ≤ Cε2 n 2 , a E0 ,0 = 0 . Let γ = E 0 − E ∞ be the energy shift. The energy shift can be bounded by |γ | ≤ Dε2 n 2 (2.10). Thus with respect to E ∞ Lemma 2.4 guarantees |z E∞ (0)| = εn + O(ε 2 n) , kη E∞ (0)kY ≤ Cε2 n 2 , a E∞ (0) ≤ Cε 2 n 2 . The second part of Proposition 4.1 thus provides the lower bound |z ∞ (t)| ≥ (1 − 2σ ) {t}−1/2 for all times t. We now prove (1.26), the last statement of Theorem 1.4. Since |γ | ≤ Dε 2 n 2 , by (2.8) we have (7.30) kQ E0 k2 = kQ E∞ k2 + 2γ (Q ∞ , R∞ ) + O(n −2 γ 2 ) . We can also estimate the energy shift γ = a∞ (0) + O(ε3 n 2 ) by Lemma 2.4 with E 1 standing for E 0 and E 2 standing for E ∞ . Thus we have Z 0 Z 0 4 9/4 a∞ (0) = B22 |z| ds + O (εn) B22 |q|4 ds + O (εn)9/4 = ∞ ∞ where q = q∞ (t). Recall B22 = + O(n ) and c1 = (Q ∞ , R∞ )−1 . On the other hand, we have dtd |q|2 = −20|q| + {t}−17/8 ; hence Z ∞ 2 |q(0)| = −20|q|4 ds + O((εn)9/4 ) . 1 c0 2 1 4 3 0 From these formulae for a∞ (0) and q(0) we have 4a∞ (0)(Q ∞ , R∞ ) + |q(0)|2 = O (εn)9/4 . Recall that the eigenfunctions for L satisfy u, v = φ1 + O(n 2 ). From the definition of z, we have |z(0)|2 = kζ (0)k2L 2 · [1 + O(n 2 )]. Since |q(0)|2 = |z(0)|2 + O((εn)9/4 ), we conclude 2γ (Q ∞ , R∞ ) = − 12 kζ (0)k2L 2 + O(ε2 n 2 (ε1/4 + n 1/4 )) and (7.31) kQ E0 k2 = kQ E∞ k2 − 12 kζ (0)k2L 2 + O ε2 n 2 (ε1/4 + n 1/4 ) . Therefore (1.26) and thus Theorem 1.4 are proven. 8 Existence of Dispersion-Dominated Solutions In this section we prove Theorem 1.5. For a given profile ξ∞ , we will construct a solution of the form (1.18), ψ = [Q(x) + a(t)R(x) + h(t, x)]ei[−Et+θ (t)] , so that ψ(t) − ψas (t) = O(t −2 ). Here a(t), θ(t) ∈ R and h(t) ∈ M. By (2.29) we have ∂t U h = −i AU h − i θ̇U h + U i −1 (F + θ̇a R) − [U, i]θ̇ h . 202 T.-P. TSAI AND H.-T. YAU (Recall U PM = U .) We single out −i θ̇U h since it is a global linear term in h. Let e h = eiθ U h; we have h = −i Ae h + eiθ U i −1 (F + θ̇a R) − [U, i]θ̇ h . (8.1) ∂t e Let e h(t) = e−i At U ξ∞ + g(t) , (8.2) h. h = U −1 e−iθ e Here g(t) ∈ X is the error term. Note that g(t) consists of both excited state and subspace of continuity components. From (8.1) we have Z t (8.3) g(t) = e−i A(t−s) eiθ U i −1 (F + θ̇a R) − [U, i]θ̇ h ds . ∞ We will solve {a, θ, g} satisfying (2.30), (2.32), and (8.3), respectively, with e h and h defined in (8.2). Note F = F(a R + h). The main term in F is ξ(t) = U −1 e−iθ (t) e−i At U ξ∞ . F0 = λQ 2|ξ |2 + ξ 2 + λ|ξ |2 ξ , By assumption ξ∞ is small in the space Z = H 2 ∩ W 2,1 (R3 ). Since U is bounded in W k, p , kU ξ∞ k H 2 ∩W 2,1 (R3 ) ≤ ε for some ε sufficiently small. From Lemmas 2.6 and 2.7, we have the bounds kξ(t)k H 2 ≤ C1 ε and kξ(t)kW 2,∞ ≤ C1 ε|t|−3/2 . Hence the term λ|ξ |2 ξ in F0 is bounded by k|ξ |2 ξ(t)k H 2 ≤ Ckξ(t)k3H 2 ≤ Cε3 . Moreover, if t > 1, 2 |ξ | ξ(t) 2 ≤ C ξ(t)2 ξ(t) ≤ Cε3 t −3 , L 2 ∞ 2 ∇ (|ξ |2 ξ )(t) 2 ≤ C ξ(t)2 ∇ 2 ξ(t) + C ξ(t) ∇ξ(t)2 ≤ Cε3 t −3 . L 2 ∞ ∞ 4 We conclude that (8.4) 2 |ξ | ξ(t) H2 ≤ Cε3 hti−3 . We also have the following estimate: For a function φ ∈ L 1 ∩ L 2 , we have (φ, |ξ |2 ξ ) ≤ min kφk L 1 · k|ξ |2 ξ k L ∞ , kφk L 2 · k|ξ |2 ξ k L 2 (8.5) ≤ C kφk L 1 ∩L 2 ε3 hti−9/2 . The other term λQ(2|ξ |2 + ξ 2 ) in F0 is actually larger than λ|ξ |2 ξ locally and can be estimated similarly. We conclude kF0 (t)k H 2 ≤ Cε2 hti−3 . If kg(t)k H 2 ≤ Cε2 hti−2 , one can prove, for example, |ξ + g|2 (ξ + g)(t) 2 ≤ Cε3 hti−3 , kQξ g(t)k H 2 ≤ Cε3 hti−7/2 . H From these estimates on F0 , the main term of F, the following bounds follow from definitions (8.3), (2.30), and (2.32) of g, a, and θ̇, respectively, (8.6) g(t) . t −2 , a . t −2 , θ̇ . t −2 , θ . t −1 . DYNAMICS OF NLS 203 One verifies with these orders that the main term of F is indeed F0 . Although θ (t) = O(t −1 ), θ only appears in the form eiθ and does not change the estimates. We now proceed to construct a solution. For convenience, we introduce a new variable ω = θ̇ . Let A be the space A = (ω, θ, a, g) : [0, ∞) → R × R × R × H 2 , |ω(t)| ≤ ε2−2σ hti−2 , |θ (t)| ≤ ε2−3σ hti−1 , |a(t)| ≤ ε2−σ hti−2 , kg(t)k H 2 ≤ ε2−σ hti−2 , where σ = 1/100 is small. We define a Cauchy sequence on the space A by iterating the following map: (cf. (2.30), (2.32), and (8.3)) ω4 (t) := − [a + (c1 R, Re F)] · [1 + a(c1 R, R) + (c1 R, Re h)]−1 , Z t 4 ω ds , θ (t) := ∞ Z t a 4 (t) := (c1 Q, Im(F + ωh)) ds , ∞ Z t 4 g (t) := e−i A(t−s) e−i A(t−s) eiθ U i −1 (F + ωa R) − [U, i]ωh ds , ∞ where F = F(a R + h) and h(t) = U −1 e−iθ (e−i At U ξ∞ + g(t)). Our initial data are ω(t) ≡ 0, θ (t) ≡ 0 , a(t) ≡ 0 , g(t) ≡ 0 . Given (ω, θ, a, g) ∈ A, using this assumption and (8.4) and (8.5), we have k|g|2 gk H 2 ≤ kgk3H 2 ≤ Cε6−3σ hti−6 , kFk H 2 ≤ Cε2 hti−3 , |(φ, F)| ≤ C kφk L 1 ∩L 2 Cε2 hti−3 , |ω4 (t)| ≤ Cε2−σ hti−2 ≤ ε2−2σ hti−2 , Z t 4 ε2−2σ hsi−2 ds ≤ ε2−3σ hti−1 , |θ (t)| ≤ ∞ Z t |a 4 (t)| ≤ Cε2 hsi−3 ds ≤ ε2−σ hti−2 , ∞ Z t 4 kg (t)k H 2 ≤ Cε2 hsi−3 ds ≤ ε2−σ hti−2 , ∞ provided that ε is small enough. In the last line we have used Lemma 2.6 on the boundedness of e−it A on X ∩ H 2 . We have also used the estimate k[U, i]hk H 2 ≤ C kξ kW 2,∞ + C kgk H 2 . 204 T.-P. TSAI AND H.-T. YAU The estimate of ξ is similar to (2.42) in Lemma 2.9, which in some sense says [U, i] is a local operator. Although it is not stated in Lemma 2.9, the proof of Lemma 2.9 in Section 9 gives the estimate. We have shown that (ω4 , θ 4 , a 4 , g 4 ) ∈ A, that is, our mapping maps A into itself. Next we show that the mapping is a contraction. Given (ω1 , θ1 , a1 , g1 ) and (ω2 , θ2 , a2 , g2 ) ∈ A, we denote δ0 = sup 8hti2 |δω(t)| + hti|δθ (t)| + 27 hti2 |δa(t)| + hti2 kδg(t)k H 2 ; 0≤t<∞ we know δ0 ≤ Cε2−3σ . Note that F0 is cancelled in δ F and that δ(eiθ ) = O(δθ). We have (the norms of F, g, and h are taken in H 2 ) kδ(|g|2 g)k H 2 ≤ C kgk2 kδgk ≤ Cε4−2σ δ0 hti−6 , |(φ, δh)| ≤ Cφ εhti−3/2 |δθ| + Cφ kδgk ≤ Cφ δ0 hti−2 , kδ Fk H 2 ≤ C ka R + hk · (εhti−3/2 |δθ| + kδgk + |δa|) ≤ Cεδ0 hti−7/2 , 5 1 |δω4 (t)| ≤ |δa| + C kδ Fk + Cεhti−3/2 |(R, δh)| ≤ δ0 hti−2 , 4 64 Z t 1 |δθ 4 (t)| ≤ |δω|ds ≤ δ0 hti−1 , 8 ∞ Z t 4 kδ Fk + ε2−2σ hsi−2 (|δω| + |(Q, δh)|)ds ≤ 2−10 δ0 hti−2 , |δa (t)| ≤ C ∞ Z t 4 kδg (t)k H 2 ≤ C ε2 hsi−3 |δθ| + kδ Fk , ∞ 1 + ε2−2σ hsi−2 (|δω| + |δa| + δ0 hsi−2 )ds ≤ δ0 hti−2 . 8 Here we have used Lemma 2.6 and the localness of [U, i] again. Therefore we have sup 8hti2 |δω4 (t)| + hti|δθ 4 (t)| + 27 hti2 |δa 4 (t)| + hti2 kδg 4 (t)k H 2 0≤t<∞ 1 ≤ δ0 , 2 and thus our map is a contraction. We conclude that we do have solutions e h with −i At the main profile e U ξ∞ . Recall ψas (t) = Q E e−i Et+iθ (t) + e−i Et et L ξ∞ . By (8.2) we have ψ(t) = [Q + a R + U −1 e−iθ (e−i At U ξ∞ + g)]e−i Et+iθ (t) = ψas (t) + [a R + U −1 e−iθ g]e−i Et+iθ (t) + e−i Et+iθ [U −1 , e−iθ ]e−i At U ξ∞ . DYNAMICS OF NLS 205 Since a(t), kg(t)k H 2 = O(t −2 ), and by the localness of [U −1 , i] [U −1 , e−iθ ]e−i At U ξ∞ = − sin θ[U, i]e−i At U ξ∞ = O(t −1 ) · O(t −3/2 ) , we have ψ(t) − ψas (t) = O(t −2 ) in H 2 . Hence Theorem 1.5 is proved. Finally, we address the remark after Theorem 1.5. Rewrite L = i(1 + E) + V1 where V1 = i −1 (V + 2λQ 2 + λQ 2 C) is a local operator. One may replace the ξ (t) with definition of ξ(t) by ξ(t) = U −1 e−iθ (t) Ue Z t i(1+E)t i(1+E)s e (8.7) ξ (t) = PL χ∞ + e(t−s)L PL χ∞ ds , ce c V1 e ∞ and proceed as in the previous proof to construct a solution h = U −1 e−iθ (Ue ξ + g). The assumption χ̂∞ (0) = 0 and ∇ χ̂∞ (0) = 0 ensures that ei(1+E)t χ∞ has a local ξ (t) decay of order O(t −7/2 ); see [5, lemma 5.2]. Hence the difference between e and ei(1+E)t χ∞ is of order O(t −5/2 ) in H 2 . Thus globally F0 (t) = O(t −3 ) and locally F0 (t) = O(t −5 ). Hence a, θ̇ = O(t −4 ) and θ(t) = O(t −3 ). Therefore in H 2 ψ(t) = [Q + U −1 e−iθ (t) Ue ξ (t)]e−i Et+iθ (t) + O(t −2 ) = Qe−i Et + ei1t χ∞ + O(t −2 ) . Here we use again that ei1t χ∞ has a fast local decay and that [U −1 , i] is a local operator. Note that (8.7) is in fact an expansion of et L WL χ∞ by the Duhamel’s formula, where WL is the wave operator of L defined in (2.36). See [5, section 5] for a similar argument. Remark (Remark on Dispersion-Dominated Solutions to Klein-Gordon Equations). We now sketch a construction for dispersion-dominated solutions to Klein-Gordon equations. We follow the notation in the introduction. For a specified profile η± , let u = ξ + g where ξ(t) = ei Bt η+ + e−i Bt η− and g denotes the rest. Then we have (∂t2 + B 2 )ξ = 0 , Hence g(t) satisfies g(t) = Z t ∞ (∂t2 + B 2 )g = λ(ξ + g)3 . ei B(t−s) − e−i B(t−s) 1 λ(ξ + g)3 ds . 2i B Since theR main source term is bounded by kξ 3 k2 ≤ kξ k2∞ kξ k2 ≤ Ct −3 , we have t kgk2 ≤ ∞ Cs −3 ds ≤ Ct −2 . Then we proceed as in Section 8 to construct one such solution by a contraction mapping argument. 9 Proof of Linear Estimates We now proceed to prove the lemmas in Section 2.4. To simplify the presentation, we will assume λ > 0. The proof for the case λ < 0 is exactly the same. Recall that X is the space of all functions in L 2 (R3 ) that are orthogonal to Q, and 5 is the orthogonal projection from L 2 (R3 ) onto X . In what follows we will only consider the restrictions of H and A on X . Hence we often omit the projection 206 T.-P. TSAI AND H.-T. YAU 5 in the definition of A. (It should be noted, however, H∗ acts on L 2 (R3 ).) In the rest of this section, when we write L 2 , W k, p , or L r2 , we often mean their intersection with X : L 2 ∩ X , W k, p ∩ X , or L r2 ∩ X . We recall assumption A2 on V . We assume that 0 is neither an eigenvalue nor a resonance for −1 + V . We also assume that V satisfies the assumption in Yajima [20] so that the W k, p estimates for k ≤ 2 for the wave operator W H holds: For a small σ > 0, |∇ α V (x)| ≤ Chxi−5−σ for |α| ≤ 2 . Also, the functions (x · ∇)k V , for k = 0, 1, 2, 3, are −1 bounded with an −1 bound less than 1: (9.1) k(x · ∇)k V φk2 ≤ σ0 k−1φk2 + C kφk2 , σ0 < 1 , k = 0, 1, 2, 3. By the assumption, the following operators are bounded in L 2 : (9.2) H∗−1/2 (x · ∇)k V H∗−1/2 , (x · ∇)k V H∗−1 , H∗−1 (x · ∇)k V , for k = 0, 1, 2, 3. Since Q is the ground state of H with V satisfying the previous assumptions, Q is a smooth function with exponential decay at infinity. Hence the above statements on V also hold for Q and Q 2 . Since V + λQ 2 and V have the same properties, in what follows we will replace V + λQ 2 in H by V and write H = H∗ + V to make the presentation simpler. So it should be kept in mind that the potential V in this section is in fact V + λQ 2 . For two operators S and T with bounded inverses, S is said to be T -bounded if ST −1 is a bounded operator. If both S and T are self-adjoint, this implies T −1 S is also bounded. A deeper result says S 1/2 is T 1/2 -bounded; see [11, theorem X.18]. We say S and T are mutually bounded if both ST −1 and T S −1 are bounded operators. This is the case if k(S − T )T −1 k(L 2 ,L 2 ) = θ < 1 for some θ. (It implies immediately that kST −1 k < 2. Since kT φk ≤ kSφk + k(T − S)φk ≤ kSφk + θ kT φk, we have kT φk ≤ C kSφk, which implies T is S-bounded.) L EMMA 9.1 For each k = 12 , 1, 32 , 2, 3, the operators H∗k , H k , and Ak are mutually bounded. P ROOF : That H∗k and H k are mutually bounded follows from our assumption on V by standard argument. To show H k and Ak are mutually bounded, it suffices to prove the cases k = 2 and k = 3 by the previous remark. We first show k(A2 − H 2 )H −2 k < 1, which implies the case k = 2. 2 (A − H 2 )H −2 = H 1/2 λQ 2 H 1/2 H −2 ≤ H 1/2 λQ 2 H −1 ≤ H∗1/2 λQ 2 H∗−1 ≤ 1/2 . The last inequality can be obtained by writing H∗1/2 Q 2 H∗−1 = H∗−1/2 Q 2 + H∗1/2 [Q 2 , H∗−1 ] = H∗−1/2 Q 2 + H∗−1/2 [Q 2 , H∗ ]H∗−1 , DYNAMICS OF NLS 207 and noting [Q 2 , H∗ ] = 1Q 2 + 2∇ Q 2 · ∇. To prove the case k = 3, it suffices to prove A6 ≤ C H 6 and H 6 ≤ C A6 . Note that ( f A6 f ) = ( f A2 A2 A2 f ) ≤ ( f A2 H 2 A2 f ) . Since A2 = H 2 + H 1/2 λQ 2 H 1/2 , we have ( f A2 H 2 A2 f ) ≤ C( f H 2 H 2 H 2 f ) + C( f (H 1/2 λQ 2 H 1/2 )H 2 (H 1/2 λQ 2 H 1/2 ) f ) where the cross terms are estimated by the Schwarz inequality. To show that the last term is bounded by C( f H 6 f ), we shall show that H 3/2 Q 2 H −5/2 is bounded in X . Rewrite H 3/2 Q 2 H −5/2 = H 3/2 H −2 Q 2 H −1/2 + H 3/2 [Q 2 , H −2 ]H −1/2 = H −1/2 Q 2 H −1/2 + H −1/2 [Q 2 , H 2 ]H −5/2 . P Since [Q 2 , H 2 ] is of the form |α|≤3 G α (x)∇ α , the operators on the right side of the equation are bounded in X . This shows A6 ≤ C H 6 . That H 6 ≤ C A6 is proven similarly. Recall the standard formula Z ∞ −σ (9.3) T = 0 1 ds , s + T sσ 0 < σ < 1. The operator T in the above formula will be A2 or H . Hence we also need to Hm estimate operators of the form s+H 2 . Clearly, for s ≥ 0, −1/2 H2 H k, p k, p ≤ C . (9.4) s + H 2 k, p k, p ≤ 1 , (W ,W ) (W ,W ) L EMMA 9.2 Let s ≥ 0. The operator H/(s + H 2 ) is bounded in W k, p ∩ X with H −1/2 . (9.5) s + H 2 k, p k, p ≤ Chsi (W ,W ) Also, for k = ±1, ±2, ±3, k 1 −k −1 hxi hxi 2 2 ≤ Chsi , s+H L ,L ( ) (9.6) k H −1/2 hxi hxi−k . 2 2 ≤ Chsi s + H2 ( L ,L ) P ROOF : We can rewrite 1 1 H = √ + √ . s + H2 H + si H − si 208 T.-P. TSAI AND H.-T. YAU 2 Therefore, to prove statements √ for H/(s + H ), it suffices to prove the corresponding statements for 1/(H ± si). We first prove (9.5) for k = 0. Let κ1 denote the eigenvalue of the excited state of H , and let P1 denote the projection onto the corresponding eigenspace. We can write 1 1 1 √ = √ P1 + W H 2 √ W ∗ PH , H + si κ1 + si p − E + si H c X where p = −i∇ and W H is the wave operator of H . Note E < 0. Since W H and W H∗ are bounded in W k, p for sufficiently nice V [20], it is sufficient√ to prove √ 2 k, p 2 that 1/( p − E ± si) are bounded in W . However, 1/( p − E ± si) are convolution operators with explicit Green functions: C −|x|(−E±√si)1/2 . e |x| √ Since |e−|x|(−E± si) | ≤ e−c|x|hsi , the L 1 norms of the Green functions are bounded by hsi−1/2 . By Young’s inequality we have 1 −1/2 , p 2 − E ± √si p p ≤ Chsi 1/4 1/2 (L ,L ) which proves (9.5) for k = 0. For k ≥ 1 and for φ ∈ W k, p , we have k/2 H k/2 H H k/2 H φ ∼ H φ = H φ 2 2 2 s+H s+H s+H p W k, p Lp L −1/2 k/2 −1/2 kφkW k, p . ≤ Chsi H φ p ∼ hsi L This proves (9.5) for k ≥ 1. For (9.6), we prove the second part. The proof for the first part is similar. For k > 0, since √ 1 1 1 k hxi , = √ √ [hxik , H + si] √ H + si H + si H + si and [hxik , H + we have √ si] = 2∇ ∗ (∇hxik ) − (1hxik ) , 1 −k hxik , hxi √ H + si k−1 1 1 ∗ −k+1 ≤C √ (∇ + 1) · hxi √ hxi H + si H + si ≤ Chsi−1/2 by induction in k. We have the same estimate for [hxik , H −1√si ], and hence (9.6) holds for positive k. The proof for the case k < 0 is similar. DYNAMICS OF NLS 209 Recall L r2 is the weighted L 2 space with norm k f k L r2 = khxir f k L 2 . L EMMA 9.3 The operators H 1/2 A−1/2 , A−1/2 H 1/2 , H −1/2 A1/2 , and A1/2 H −1/2 are bounded operators in W k, p ∩ X and L r2 ∩ X . P ROOF : By (9.3) we can write Z ∞ 1 ds 1/2 −1/2 1/2 H A =H 2 1/2 2 1/2 s + H + H λQ H s 1/4 Z0 ∞ 1 1 = H 1/2 + H 1/2 λQ 2 2 s + H s + H 0 j ∞ X H ds 1 1/2 Q Q H λQ 2 2 s 1/4 s + H s + H j=0 j Z ∞ ∞ X H H H −1/2 ds =1+ λQ Q Q H . λQ s + H2 s + H2 s + H2 s 1/4 0 j=0 H −1/2 by Lemma 9.2, we have Since k s+H 2 k ≤ hsi Z 1/2 −1/2 H A (W k, p ,W k, p ) ≤ 1 + C ∞ 0 hsi−1/2 n ∞ X ds (n 2 hsi−1/2 ) j nhsi−1/2 1/4 s j=0 ≤ 1 + Cn 2 . Similarly, −1/2 1/2 A H 2 (W k, p ,W k, p ) ≤ 1 + Cn . Also, using (9.6), for r ≤ 3 we have 1/2 −1/2 H A 2 2 + A−1/2 H 1/2 2 2 ≤ 1 + Cn 2 . ( L ,L ) ( L ,L ) r r r r The above proves that H 1/2 A−1/2 and A−1/2 H 1/2 are bounded in W k, p and L r2 . Indeed, we have proven (9.7) 3 1/2 −1/2 hxi (H A − 1)hxi3 ( L 2 ,L 2 ) + hxi3 (A−1/2 H 1/2 − 1)hxi3 ( L 2 ,L 2 ) ≤ Cn 2 . We now consider H −1/2 A1/2 and A1/2 H −1/2 . Since Z ∞ 1 ds 1/2 2 −3/2 2 A =A A =A , 2 s + A s 3/4 0 210 T.-P. TSAI AND H.-T. YAU we have H −1/2 A 1/2 Z ∞ 1 ds λQ H ) 2 + H 1/2 λQ 2 H 1/2 s 3/4 s + H 0 Z ∞ 1 ds = (H 3/2 + λQ 2 H 1/2 ) 2 1/2 2 1/2 s + H + H λQ H s 3/4 0 =H −1/2 (H + H 2 1/2 2 1/2 = I1 + λQ 2 I2 . The main term is I1 . The term I2 is similar to H 1/2 A−1/2 , and its integrand has a better decay in s for large s. Hence 2 λQ I2 ≤ Cλ kI2 k ≤ Cn 2 . For the main term I1 , Z ∞ I1 = 1 + 0 j ∞ X H H2 ds H Q Q Q Q H −1/2 3/4 . 2 2 2 s+H s+H s+H s j=0 Hence Z ∞ kI1 k ≤ 1 + C n Z 0 ∞ ≤1+C 0 ∞ X ds (n 2 hsi−1/2 ) j nhsi−1/2 3/4 s j=0 n 2 hsi−1/2 ds ≤ 1 + Cn 2 . s 3/4 Here the norms are taken in (W k, p , W k, p ) and (L r2 , L r2 ). Hence we have proven Lemma 9.3. In fact, the last part of the above proof also shows (9.8) hxi3 (H −1/2 A1/2 − 1)hxi3 ( L 2 ,L 2 ) + hxi3 (A1/2 H −1/2 − 1)hxi3 ( L 2 ,L 2 ) ≤ Cn 2 . P ROOF OF L EMMA 2.9: The above lemma proves that U0 and U0−1 are bounded in W k, p and L r2 . Moreover, (9.7) and (9.8) mean that U0 −1 and U0−1 −1 are “local” operators. In particular, they imply k[U0 , i] φk L 8/7 ∩L 4/3 ≤ C kφk L 4 . Since P1 0 e e U = U 5 where 5 = 0 5 is a bounded projection, U is also bounded. Similarly U −1 is also bounded. Moreover, we have 0 1 e , i] + [U0 , i]5 e = U0 (P1 − 5) e, [U, i] = U0 [5 + [U0 , i]5 1 0 where by (2.18) P1 − 5 = −c1 |5Ri hQ| is a local operator; hence [U, i] satisfies the last estimate in Lemma 2.9. Lemma 2.9 is thus proven. DYNAMICS OF NLS 211 P ROOF OF L EMMA 2.10: We need only to prove the statement on W A . Note that the estimates (9.7) and (9.8) also show, for any φ ∈ L 2 , (U ±1 − 1)e−it H∗ φ → 0 (9.9) in L 2 as t → ∞ . Notice W A = lim eit A e−it H∗ = lim U et LU −1 e−it H∗ t→∞ t→∞ = lim U et L e−it H∗ + lim U et L (U −1 − 1)e−it H∗ . t→∞ By (9.9) we have t→∞ W A = lim U et L e−it H∗ = U WL . t→∞ The boundedness of W A follows from that of U and WL . This proves Lemma 2.10. P ROOF OF L EMMA 2.7: Since e−it A PcA φ = W A e−it H∗ W A∗ PcA φ , the estimate (2.39) follows from the usual (L p , L q ) estimate for e−it H∗ and the boundedness of W A and PcA in L p spaces. To prove (2.40), either we prove the boundedness of W A in weighted spaces L r2 , or we use the Mourre estimate. We will follow the second approach and the argument in [16]. Let a = 2κ. We consider intervals 1 = (a − r, a + r ). Let g1 (t) = g0 ((t − a)/r ), where g0 is a fixed smooth function with support in (−2, 2) and g0 (t) = 1 for |t| < 1. We will consider g1 (A) with r small enough. Let D = x p + px, p = −i∇, and the commutators ad0D (A) = A , k adk+1 D (A) = [ad D (A), D] . We need to prove the following lemma: L EMMA 9.4 For 1 small enough, the Mourre estimate g1 (A)[i A, D]g1 (A) ≥ θg1 (A)2 holds for some θ > 0. Also, g1 (A) adkD (A)g1 (A) are bounded operators in L 2 for k = 0, 1, 2, 3. We will use the following lemma: L EMMA 9.5 The operators H −3 D k H m/2 hxi−3 and hxi−3 H m/2 D k H −3 are bounded in L 2 for k, m = 0, 1, 2, 3. P ROOF : This is standard and we only sketch the proof. If m is even, we can compute the commutator [D k , H m/2 ] explicitly and estimate H −3 D k H m/2 hxi−3 = H −3 H m/2 D k hxi−3 + H −3 [D k , H m/2 ]hxi−3 . 212 T.-P. TSAI AND H.-T. YAU If m is odd, we write H −3 k D H m/2 hxi −3 Z = ∞ H −3 D k H (m+1)/2 0 1 ds hxi−3 √ s+H s and proceed as in the case when m is even, by using (9.6). Here we have used formula (9.3). P ROOF OF L EMMA 9.4: Let G = A − H and write A = H + G. Since [i A, D] = [H∗ + V + G, i D] = −1 + [V + G, i D] = A − V − G + [V + G, i D] and g1 (A)Ag1 (A) ≥ 2θg1 (A)2 for some 2θ > 0, it suffices to show that, for M = −V + [V, i D], −G, and [G, D], the operators g1 (A)Mg1 (A) = (g1 (A)H∗2 )(H∗−2 M H∗−2 ) (H∗2 g1 (A)) are bounded by g1 (A)2 , and the bound goes to zero when the interval 1 shrinks to zero. Since both g1 (A)H∗2 = (g1 (A)A2 )(A−2 H∗2 ) and H∗2 g1 (A) = (H∗2 A−2 )(A2 g1 (A)) are bounded and converge to zero weakly when 1 shrinks to zero, this will be true if one can show that H∗−2 M H∗−2 is compact. The case M = −V + [V, i D] is standard and follows from our assumption, so we only consider H∗−2 G H∗−2 and H∗−2 [G, D]H∗−2 . We proceed to find an explicit form of G. By (9.3) with T = A2 , σ = 12 , we write Z ∞ 1 ds −1 A = √ 2 1/2 2 1/2 s + H + H λQ H s 0 j Z ∞ ∞ X 1 H 1 1 ds 1/2 + H λQ Q Q H 1/2 = λQ √ 2 2 2 2 s+H s+H s+H s+H s 0 j=0 = H −1 + H −1/2 hxi−3 J0 hxi−3 H −1/2 where Z J0 = ∞ 0 j ∞ X H H H ds hxi λQ Q Q hxi3 √ . λQ 2 2 s + H2 s + H s + H s j=0 3 By Lemma 9.2, Z kJ0 k(L 2 ,L 2 ) ≤ 0 ∞ hsi−1/2 · n 2 · hsi−1/2 s −1/2 ds ≤ Cn 2 . DYNAMICS OF NLS 213 Hence A = A2 A−1 (9.10) = (H 2 + H 1/2 λQ 2 H 1/2 ) H −1 + H −1/2 hxi−3 J0 hxi−3 H −1/2 = H +G, G = H 1/2 λQ 2 H −1/2 + H 3/2 hxi−3 J0 hxi−3 H −1/2 + H 1/2 λQ 2 hxi−3 J0 hxi−3 H −1/2 . Since H −1/2 hxi−1 and hxi−1 H −1/2 are compact, from (9.10) H∗−2 G H∗−2 is compact. We can also write H∗−2 G D H∗−2 = H∗−2 G H 1/2 hxi · hxi−1 H −1/2 D H∗−2 . The second operator is bounded by Lemma 9.5. The first is compact since its terms are of the form H −m · hxi−k · (bounded operator). Similarly, H∗−2 DG H∗−2 is also compact. Hence we conclude the Mourre estimate. To show that g1 (A) adkD (A)g1 (A) are bounded for k = 0, 1, 2, 3, we rewrite g1 (A) adkD (A)g1 (A) = (g1 (A)A3 )(A−3 H 3 )(H −3 adkD (A)H −3 )(H 3 A−3 )(A3 g1 (A)) . We need only to show that H −3 adkD (A)H −3 are bounded since the other terms are bounded by Lemma 9.1. Recall A = H + G. It is standard to prove that H −3 adkD (H )H −3 is bounded. For H −3 adkD (G)H −3 , since it is a sum of terms of the form H −3 D k G D m H −3 , k + m ≤ 3 , it suffices to show that these terms are bounded. By the explicit form (9.10) of G and Lemma 9.5, they are indeed bounded. For example, H −3 D 2 H 3/2 hxi−3 J0 hxi−3 H −1/2 D 1 H −3 = −3 2 3/2 −3 −3 −1/2 1 −3 H D H hxi D H J0 hxi H , a product of three bounded operators. We conclude that g1 (A) adkD (A)g1 (A) are bounded for k = 0, 1, 2, 3. With Lemma 9.4 (cf. the remark in [16, p. 27]), the minimal velocity estimate in [6] and theorem 2.4 of [14] implies F(D ≤ θt/2)e−i At g1 (A) hDi−3/2 2 2 ≤ Chti−5/4 . (L ,L ) The same argument in [16] then gives the desired decay estimate (2.40). P ROOF OF L EMMA 2.8: Let ψn = PcA 5φ0 φ12 and ψ0 = PcH1 φ0 φ12 . Recall H1 = −1 + V − e0 . We have ψn = ψ0 + O(n 2 ). We write ψn = ψ0 + bφ1 + η, where 214 T.-P. TSAI AND H.-T. YAU η ∈ Hc (H1 ) and b, η = O(n 2 ). Rewrite 1 ψn ψn , Im A − 0i − 2κ Z ∞ = Im i ψn , e−it (A−0i−2κ) ψn dt 0 (9.11) Z = Im i ∞ ψn , e−it (H1 −0i−2κ) ψn dt 0 (9.12) Z + Im i 0 ∞ Z t ψn , e−i(t−s)(A−0i−2κ) (λQ 2 + G)e−is(H1 −0i−2κ) ψn ds dt . 0 The main term lies in (9.11). It is Z ∞ −it (H1 −0i−2κ) Im i ψ0 dt = ψ0 , Im ψ0 , e 0 1 ψ0 H1 − 0i − 2κ , which is the desired main term in Lemma 2.8. We want to show that the rest of (9.11) and (9.12) are integrable and of order O(n 2 ). Recall that we write ψn = ψ0 + bφ1 + η. For the term η in ψn , by the decay estimate we have (9.13) ψn , e−it (H1 −0i−2κ) η ≤ Chti−3/2 kψn k L 1 ∩L 2 kηk L 1 ∩L 2 ≤ Chti−3/2 n 2 ; hence this term is integrable. Also, since H1 φ1 = e01 φ1 , ψn , e−it (H1 −0i−2κ) bφ1 = ψn , e−it (e01 −2κ−0i) bφ1 , so we can integrate this oscillation term explicitly. (The boundary term at t = ∞ vanishes due to the decay of e−it (−0i) .) We conclude that the rest of (9.11) is integrable and of order O(n 2 ). For (9.12), it suffices to show its integrability since λQ 2 + G gives the order O(n 2 ). Rewrite the last ψn in (9.12) as bφ1 + PcH1 ψn . For the part containing bφ1 , we have ψn , e−i(t−s)(A−0i−2κ) (λQ 2 + G)e−is(H1 −0i−2κ) bφ1 = ψn , e−it (A−0i−2κ) eis(A−e01 ) (λQ 2 + G)bφ1 . Integration in s gives (A − e01 )−1 ψn , e−it (A−0i−2κ) (λQ 2 + G)bφ1 . Since e01 lies outside the continuous spectrum of A, the last expression is integrable in t following the same argument as (9.13). For the part containing PcH1 ψn , since (λQ 2 + G) is a “local” operator in the sense that it sends L ∞ functions to L 1 , we DYNAMICS OF NLS 215 have ψn , e−i(t−s)(A−0i−2κ) (λQ 2 + G)e−is(H1 −0i−2κ) P H1 ψn ≤ c Cn 2 ht − si−3/2 hsi−3/2 kψn k2L 1 ∩L 2 , which can be integrated in s and t. Hence we have proven Lemma 2.8. P ROOF OF L EMMA 2.6: Given φ ∈ H k ∩ X , k = 0, 1, 2, let u(t) = e−it A φ ∈ X . We have dtd (u(t), Am u(t)) = 0; hence (u(t), Am u(t)) is a conserved quantity. When k = 0, this implies that the L 2 norm is conserved. For k = 2, we have (u, A2 u) = u, (H 2 + H 1/2 λQ 2 H 1/2 )u = (u, H 2 u) + O(n 2 kuk2H 1 ) ≤ C kuk2H 2 . On the other hand, since kuk2H 1 ≤ C kuk2H 2 + C kuk2L 2 and (u, u) ≤ C(u, A2 u) due to the spectral gap of A, we have C −1 kuk2H 2 ≤ (u, H 2 u) = (1 − O(n 2 ))−1 (u, A2 u) + O(n 2 ) kuk2L 2 ≤ C(u, A2 u) . Thus we have (u(t), A2 u(t)) ≈ ku(t)k2H 2 in the sense of (1.23), and we have ku(t)k2H 2 ≈ (u(t), A2 u(t)) = (u(0), A2 u(0)) ≈ ku(0)k2H 2 . The case k = 2 is thus proven. The case k = 1 can be obtained by interpolation. Acknowledgments. It is with great pleasure that we thank M. I. Weinstein for explaining the beautiful ideas behind his work with A. Soffer [16]. We also thank S. Cuccagna for his very helpful comments and, in particular, for pointing out a gap in the proof of Theorem 1.5 in the manuscript. The work of H.-T. Yau was partially supported by NSF Grant DMS-0072098. Part of this work was done when both authors were visiting the Center for Theoretical Sciences, Taiwan, in summer 2000. Bibliography [1] Bourgain, J.; Wang, W. Construction of blowup solutions for the nonlinear Schrödinger equation with critical nonlinearity. Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4) 25 (1997), no. 1-2, 197–215 (1998). [2] Bronski, J. C.; Jerrard, R. L. Soliton dynamics in a potential. Math. Res. Lett. 7 (2000), no. 2-3, 329–342. [3] Buslaev, V. S.; Perel0 man, G. S. On the stability of solitary waves for nonlinear Schrödinger equations. Nonlinear evolution equations, 75–98. American Mathematical Society Translations, Series 2, 164. American Mathematical Society, Providence, R.I., 1995. [4] Cuccagna, S. Stabilization of solutions to nonlinear Schrödinger equations. Comm. Pure Appl. Math. 54 (2001), no. 9, 1110–1145. [5] Fröhlich, J.; Tsai, T.-P.; Yau, H.-T. On the point-particle (Newtonian) limit of the non-linear Hartree equation. Comm. Math. Phys., in press. 216 T.-P. TSAI AND H.-T. YAU [6] Hunziker, W.; Sigal, I. M.; Soffer, A. Minimal escape velocities. Comm. Partial Differential Equations 24 (1999), no. 11-12, 2279–2295. [7] Jensen, A.; Kato, T. Spectral properties of Schrödinger operators and time-decay of the wave functions. Duke Math. J. 46 (1979), no. 3, 583–611. [8] Journe, J.-L.; Soffer A.; Sogge, C. D. Decay estimates for Schrödinger operators. Comm. Pure Appl. Math. 44 (1991), no. 5, 573–604. [9] Pillet, C.-A.; Wayne, C. Invariant manifolds for a class of dispersive, Hamiltonian, partial differential equations. J. Differential Equations 141 (1997), no. 2, 310–326. [10] Rauch, J. Local decay of scattering solutions to Schrödinger’s equation. Comm. Math. Phys. 61 (1978), no. 2, 149–168. [11] Reed, M.; Simon, B. Methods of modern mathematical physics. II. Fourier analysis, selfadjointness. Academic, New York–London, 1975. [12] Reed, M.; Simon, B. Methods of modern mathematical physics. III. Scattering theory. Academic, New York–London, 1979. [13] Rose, H. A.; Weinstein, M. I. On the bound states of the nonlinear Schrödinger equation with a linear potential. Phys. D 30 (1988), no. 1-2, 207–218. [14] Skibsted, E. Propagation estimates for N -body Schroedinger operators. Comm. Math. Phys. 142 (1991), no. 1, 67–98. [15] Soffer, A.; Weinstein, M. I. Multichannel nonlinear scattering for nonintegrable equations. I. II. The case of anisotropic potentials and data. Comm. Math. Phys. 133 (1990), no. 1, 119–146; J. Differential Equations 98 (1992), no. 2, 376–390. [16] Soffer, A.; Weinstein, M. I. Resonances, radiation damping and instability in Hamiltonian nonlinear wave equations. Invent. Math. 136 (1999), no. 1, 9–74. [17] Tsai,T.-P.; Yau, H.-T. Relaxation of excited states in nonlinear Schrödinger equations. Submitted. [18] Weinstein, M. I. Modulational stability of ground states of nonlinear Schrödinger equations. SIAM J. Math. Anal. 16 (1985), no. 3, 472–491. [19] Weinstein, M. I. Lyapunov stability of ground states of nonlinear dispersive evolution equations. Comm. Pure Appl. Math. 39 (1986), no. 1, 51–68. [20] Yajima, K. The W k, p -continuity of wave operators for Schrödinger operators. J. Math. Soc. Japan 47 (1995), no. 3, 551–581. TAI -P ENG T SAI Institute for Advanced Study Einstein Drive Princeton, NJ 08540 E-mail: ttsai@ias.edu Received November 2000. Revised July 2001. H ORNG -T ZER YAU Courant Institute 251 Mercer Street New York, NY 10012 E-mail: yau@cims.nyu.edu