ebook img

Spectral gaps in Wasserstein distances and the 2D stochastic Navier--Stokes equations PDF

0.45 MB·English
Save to my drive
Quick download
Download
Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.

Preview Spectral gaps in Wasserstein distances and the 2D stochastic Navier--Stokes equations

TheAnnalsofProbability 2008,Vol.36,No.6,2050–2091 DOI:10.1214/08-AOP392 (cid:13)c InstituteofMathematicalStatistics,2008 SPECTRAL GAPS IN WASSERSTEIN DISTANCES AND THE 2D 9 0 STOCHASTIC NAVIER–STOKES EQUATIONS 0 2 By Martin Hairer1 and Jonathan C. Mattingly2 n a University of Warwick and Duke University J 2 We develop a general method to prove the existence of spectral 2 gaps for Markov semigroups on Banach spaces. Unlike most previ- ous work, the type of norm we consider for this analysis is neither ] a weighted supremum norm nor an L p-type norm, but involves the R derivativeoftheobservableaswellandhencecanbeseenasatypeof P 1-Wasserstein distance. This turns out to be a suitable approach for . h infinite-dimensional spaces where theusual Harris or Doeblin condi- t tions,whicharegeared towardtotalvariationconvergence,oftenfail a to hold. In the first part of this paper, we consider semigroups that m haveuniformbehaviorwhichonecanviewastheanalogofDoeblin’s [ condition. We then proceed to study situations where the behavior 2 is not so uniform, but the system has a suitable Lyapunov struc- v ture, leading to a typeof Harris condition. Wefinally show that the 9 latterconditionissatisfiedbythetwo-dimensionalstochasticNavier– 7 Stokes equations, even in situations where the forcing is extremely 4 degenerate.Usingtheconvergenceresult,weshowthatthestochastic 2 Navier–Stokesequations’invariantmeasuresdependcontinuouslyon 0 theviscosity and thestructureof theforcing. 6 0 / h 1. Introduction. This work is motivated by the study of the two-dimen- t a sional stochastic Navier–Stokes equations on the torus. However, the results m andtechniques aremoregeneral.Themainabstractresultofthepapergives : a criterion guaranteeing that a Markov semigroup on a Banach space has a v i spectral gap in a particular 1-Wasserstein distance. (In the sequel, we will X simply write Wasserstein for 1-Wasserstein.) To the best of our knowledge, r a Received May 2006; revised May 2007. 1Supportedin part by EPSRCAdvanced Research Fellowship Grant EP/D071593/1. 2Supported in part by NSF PECASE Award DMS-04-49910 and an Alfred P. Sloan foundation fellowship. AMS 2000 subject classifications. 37A30, 37A25, 60H15. Key words and phrases. Stochastic PDEs, Wasserstein distance, ergodicity, mixing, spectral gap. This is an electronic reprint of the original article published by the Institute of Mathematical Statistics in The Annals of Probability, 2008,Vol. 36, No. 6, 2050–2091. This reprint differs from the original in pagination and typographic detail. 1 2 M. HAIRERANDJ. MATTINGLY these results are thefirstresults providinga spectral gap in this, or any sim- ilar, setting. Inturn,the existence of a spectral gap implies that theMarkov semigroup possesses a unique, exponentially mixing invariant measure. Theresultsofthisarticlerelyontheexistenceofa“gradientestimate”in- troduced in [21] in the study of the degenerately forced Navier–Stokes equa- tions on the two-dimensional torus. This estimate was used there in order to show that the corresponding Markov semigroup satisfies the “asymptotic strong Feller” property, also introduced in [21]. In this work, we show that gradient estimates of this form not only are usefulto show uniquenessof the invariant measure, but are an essential ingredient to obtain the existence of a spectral gap for a large class of systems. In this introductory section, we concentrate on the two-dimensional stochastic Navier–Stokes equations on a torus to show how the main results can be applied. At the end of this section, we give an overview of the paper. Recall that the Navier–Stokes equations describing the evolution of the velocity field v(x,t) (with x T2) of a fluid under the influence of a body force F¯(x)+F(x,t) are given∈by (SNS) ∂ v=ν∆v (v )v p+F¯+F, divv=0, t − ·∇ −∇ where the pressure p(x,t) is determined by the algebraic condition divv=0. We consider for F a Gaussian stochastic forcing, that is, centered, white in time, colored in space and such that F¯(x)dx= F(x)dx=0. Since the gradient part of the forcing is canceleRd by the prRessure term, we assume without loss of generality that divF¯=divF =0. More precisely, we assume that for i,j 1,2 ∈{ } EF (x,t)F (x′,t′)=δ(t t′)Q (x x′), i j ij − − 2 ∂2Q =0, Q (x)dx=0. ij ij ij Z iX,j=1 AlthoughweareconfidentthatourresultsarevalidforQsufficientlysmooth, we restrict ourselves to the case where Q is a trigonometric polynomial, so that we can make use of the bounds obtained in [21, 39]. Instead of considering the velocity (SNS) directly, we will consider the equation for the vorticity w= v=∂ v ∂ v . Note that v is uniquely 1 2 2 1 ∇∧ − determined from w (we will write v= w) through the conditions K w= v, divv=0, v(x)dx=0. ∇∧ Z When written in terms of w, (SNS) is equivalent to (1) ∂ w=ν∆w ( w) w+f¯+f, t − K ·∇ SPECTRAL GAPS IN WASSERSTEINDISTANCES 3 where we have defined f = F and f¯= F¯. Note that since f is ∇∧ ∇∧ translation invariant, one can write it as f(x,t)=Re q eikxξ (t), k k k∈ZX2\{0} where the ξ are independent white noises and where q = q . We can k k −k therefore identify the correlation function Q with a vector q in ℓ2, the set of + square integrable sequences with positive entries. Denoting by the set of Z indices k for which q =0,wewillmake throughoutthisarticle thefollowing k 6 assumptions: Assumption 1. Only finitely many of the q ’s are nonzero and f¯ lies k in the span of eikx q =0 . Furthermore, generates Z2 and there exist k { | 6 } Z k,ℓ with k = ℓ . ∈Z | |6 | | Remark1.1. Theassumptionthatonlyafinitenumberofq arenonzero k isonlyatechnicalassumptionreflectingadeficiencyin[39].Alloftheresults of this article certainly hold if the first part of Assumption 1 is replaced by an appropriate decay property for the q . Note, for example, that in [21], k Section 4.5, it is shown that there exists an N such that if the range of Q ∗ contains eikx k <N , f¯is as in Assumption 1, and q2< , then all of { || | ∗} k ∞ the results of this paper hold. In particular, this allowPs infinitely many q k to be nonzero. Remark1.2. Usingtheresultsin[3]onecanremovetherestrictionthat theforcingneedconsistofFouriermodesandreplaceitwiththerequirement that the forced functions span the Fourier modes required above. Since this is not the main point of this article, we do not elaborate further here. Remark1.3. Itisclearthattheassumptionthatf¯ span eikx q =0 k is far from optimal. The correct result likely places no re∈stricti{on on|f¯o6the}r thanitbesufficientlysmooth.Thismoredelicateresultrequiresanimproved understanding of the control problem obtained by replacing the noise by controls. Some steps in this direction have been made [1, 2, 47], but the current results are not sufficient for our needs. Nonetheless, the present assumption on f¯seems reasonable from a modeling perspective where one wouldlikely havesomenoiseinallof thedirections onwhichthebodyforces act. We will consider (SNS) as an evolution equation in the subspace of H H1 that consists of velocity fields v with divv =0 in the sense of distri- butions. Note that this is equivalent to w L 2. We make a slight abuse of ∈ 4 M. HAIRERANDJ. MATTINGLY notation and denote by the transition probabilities for (SNS), as well as t P the corresponding Markov semigroup on , that is, H (v,A)=P(v(t, ) A v(0, )=v), t P · ∈ | · for every Borel set A , and ⊂H ( φ)(v)= φ(v′) (v,dv′), ( ∗µ)(A)= (v,A)µ(dv) Pt Z Pt Pt Z Pt H H for every φ: R and probability measure ν on . Analogously we define H→ H the projection µφ= φ(x)µ(dx). It was shown in [21] that Assumption 1 implies that (SNS) aRdmits a unique invariant measure µ , that is, µ is a ⋆ ⋆ probability measure on such that ∗µ =µ for every t 0. H Pt ⋆ ⋆ ≥ This article is concerned with whether, for an arbitrary probability mea- sureµ, ∗µ µ (as t )andinwhatsensethisconvergencetakes place. Pt → ⋆ →∞ Note that (SNS) is not expected to have the strong Feller property, so that it is a fortiori not expected that ∗µ µ in the total variation topology Pt → ⋆ if µ and µ are mutually singular. (See [9] for a general discussion of the ⋆ strong Feller property in infinite dimensions and [21] for a discussion of its limitations in the present setting.) In order to state the main result of the present article, we introduce the following norm on the space of smooth observables φ: R: H→ φ =supe−ηkvk2(φ(v) + Dφ(v) ). η k k | | k k v∈H Here, we denoted by Dφ the Fr´echet derivative of φ. With this notation, we will show that the operator has a spectral gap in the norm in the t η P k·k following sense: Theorem 1.4. Consider (SNS) in the range of parameters allowed by Assumption 1. For every η small enough there exist constants C and γ such that φ µ φ Ce−γt φ , t ⋆ η η kP − k ≤ k k for every Fr´echet differentiable function φ: R and every t 0. H→ ≥ In[19]asimilar operator-normestimate on was obtained in aweighted t P total variation norm ( without the Dφ term) when the forcing was η k·k k k spatially rough and nondegenerate. Our setting is quite different. The spa- tially rough and nondegenerate forcing makes the analysis much closer to the finite-dimensional setting. It is not expected that such estimates hold in the total variation norm in the setting of this article. We should also remark that previous estimates, such as [20, 29, 37, 38], giving simply expo- nential convergence to equilibrium are weaker and the results in this article represent a significant and new extension of those results. SPECTRAL GAPS IN WASSERSTEINDISTANCES 5 It is sometimes of interest to know that the structure functions of the so- lution to (SNS) converge to the structure functions determined by µ . This ⋆ is not an immediate consequence of Theorem 1.4 because point evaluations of thevelocity field arenotcontinuous functions on .Thesmoothingprop- H erties of (SNS) nevertheless allow us to show the following result, which is an immediate consequence of Theorem 1.4 and Proposition 5.12 below. Theorem 1.5. Consider (SNS) in the range of parameters allowed by Assumption 1. Let n 1 and define the n-point structure functions by ≥ n(x ,...,x )= v(x ) v(x )µ (dv). 1 n 1 n ⋆ S Z ··· Then, for every η>0, there exist constants C and γ>0 such that, for every v , one has the bound 0 ∈H n sup E v(x ,t) n(x ,...,x ) Ceηkv0k2−γt, (cid:12) i 1 n (cid:12) x1,...,xn(cid:12)(cid:12) iY=1 −S (cid:12)(cid:12)≤ (cid:12) (cid:12) for every t>1. He(cid:12)re, v(x,t) is the solution of ((cid:12)SNS) with initial condition v . 0 The remainder of this article is structured as follows. In Section 2, we begin with an abstract discussion of our ideas in a setting with uniform estimates. In Section 3, we give the main theoretical results of the paper whichcombinetheideasfromthefirstsectionwithestimatesstemmingfrom an assumed Lyapunov structure. The convergence is measured in a distance in which paths are weighted by the Lyapunov function. We then turn in Section 5 to the specifics of the stochastic Navier–Stokes equation and show that it satisfies the general theorems from Section 3. In Section 5.3, we show that the Markov semigroup generated by (SNS) is strongly continuous on a suitable Banach space and that its generator has a spectral gap there. Then in Section 5.5, we demonstrate the power of the spectral gap estimates by giving a short proof that (SNS)’s invariant measures depend continuously on all the parameters of the equation. 2. A simplified, uniform setting. In this section, we illustrate many of the main ideas used throughout this article in a simplified setting. We con- sider the analogue of one of the simplest (and yet powerful) conditions for a Markov chain with transition probabilities to have a unique invariant P measure, namely Doeblin’s condition3: 3Doeblin’s original condition was the existence of a probability measure ν and a con- stant ε>0 such that P(x,A)≥ε whenever ν(A)>1−ε; see [11, 12]. It turns out that, provided that the Markov chain is aperiodic and ψ-irreducible, this is equivalent to the (in general stronger) condition given here; see [40], Theorem 16.0.2. 6 M. HAIRERANDJ. MATTINGLY Theorem2.1 (Doeblin). Assume that there exist δ>0 anda probability measure ν such that (x, ) δν for every x. Then, there exists a unique P · ≥ probability measure µ such that ∗µ =µ . Furthermore, one has φ ⋆ ⋆ ⋆ P kP − µ φ (1 δ) φ µ φ for every bounded measurable function φ. ⋆ ∞ ⋆ ∞ k ≤ − k − k Atypical exampleofasemigroupforwhich Theorem2.1can beappliedis given byanondegeneratediffusiononasmoothcompact manifold.Theorem 2.1showsthefundamentalmechanismforconvergencetoequilibriumintotal variation norm. It is simple because the assumed estimates are extremely uniform. In this section we give a theorem guaranteeing convergence in a Wasserstein distance with assumptions analogous to Doeblin’s condition. A classical generalization of Doeblin’s condition was made by Harris [22] who showed how to combine the existence of a Lyapunov function and a Doeblin-like estimate localized to a sufficiently large compact set to prove convergencetoequilibrium.Wewillgivea“Harris-like”versionofourresults in Section 3. 2.1. Spectral gap under uniform estimates. The aim of this section is to present avery simplecondition that ensuresthat a Markov semigroup on t P a Banach space yields a contraction operator on the space of probability H measures endowed with a Wasserstein distance. One can view it as a ver- sion of Doeblin’s condition for the Wasserstein distance instead of the total variation distance. The main motivation for using a distance that metrizes the topology of weak convergence is that probability measures on infinite- dimensional spaces tend to bemutually singular, so that strong convergence is not expected to hold in general, even for ergodic systems. The first assumption captures the regularizing effect of the Markov semi- group. While it does not imply that one function space is mapped into a more regular one as often occurs, it does say that at least gradients are decreased. Assumption2. Thereexistconstants α (0,1),C >0andT >0such 1 1 ∈ that (2) D φ C φ +α Dφ , t ∞ ∞ 1 ∞ k P k ≤ k k k k for every t T and every Fr´echet differentiable function φ: R. 1 ≥ H→ Remark2.2. Atypicalwayofchecking(2)istofirstshowthatforevery t 0, is boundedas an operator on the space with norm φ + Dφ . t ∞ ∞ ≥ P k k k k It then suffices to check that (2) holds with α <1 for one particular time t 1 to deduce from the semigroup property that D φ C( φ +e−γt Dφ ) t ∞ ∞ ∞ k P k ≤ k k k k is valid with some γ>0 for every t 0. ≥ SPECTRAL GAPS IN WASSERSTEINDISTANCES 7 Remark2.3. Ifthespace isactuallyacompactmanifold,(2)together H with the Arzel`a–Ascoli theorem implies that the essential spectral radius of (as an operator on the space with norm φ + Dφ ) is strictly t ∞ ∞ P k k k k smaller than 1. This is a well-known and often exploited fact4 in the theory of dynamical systems. A bound like (2) is usually referred to as the Lasota– Yorke inequality [30, 31] or the Ionescu–Tulcea–Marinescu inequality [26] andisusedtoshowtheexistenceofabsolutelycontinuousinvariantmeasures when is the transfer operator acting on densities. Usually, it is used with t P two different H¨older norms on the right-hand side. The present application with a Lipschitz norm and an infinity norm has a different flavor. This bound alone is of course not enough in general to guarantee the uniqueness of the invariant measure. (Counterexamples with =S1, the H unitcircle,caneasilybeconstructed.)Furthermore,wearemainlyinterested in the case where is not even locally compact. H Inordertoformulateoursecondassumption,weusethenotation (µ µ ) 1 2 C for the set of all measures Γ on such that Γ(A )=µ (A) and 1 H×H ×H Γ( A) = µ (A) for every Borel set A . Such a measure Γ on the 2 H × ⊂ H product space is referred to as a coupling of µ and µ . We also denote 1 2 by ∗ the semigroup acting on probability measures which is dual to . Pt Pt With these notation, our second assumption, which is a form of uniform topological irreducibility, reads: Assumption 3. For every δ >0, there exists a T =T (δ) so that for 2 2 any t T there exists an a>0 so that 2 ≥ sup Γ (x′,y′) : x′ y′ δ a, Γ∈C(Pt∗δx,Pt∗δy) { ∈H×H k − k≤ }≥ for every x,y . ∈H Remark 2.4. Recall that the total variation distance between proba- bility measures can be characterized as one minus the supremum over all couplings of the mass of the diagonal. Therefore, if we set δ=0 in Assump- tion 3, we get (x, ) (y, ) 1 a for every x and y. By [40], t t TV kP · −P · k ≤ − Theorem 16.0.2, this is equivalent to the assumption of Theorem 2.1, so that the results in this section can be viewed as an analogue of Doeblin’s theorem. 4It can be obtained as a corollary of the fact that the essential spectral radius of an operator T can be characterized as the supremum over all λ>0 such that there exists a singular sequence {f } with kf k=1 and kTf k≥λ for every n. A slightly different n n≥0 n n proof can be found in [23] and is directly based on the study of the essential spectral radiusbyNussbaum[42].Itis,however,verycloseinspirittothemuchearlierpaper[26]. 8 M. HAIRERANDJ. MATTINGLY To measure the convergence to equilibrium, we will use the following distance function on : H (3) d(x,y)=min 1,δ−1 x y , { k − k} where δ is a small parameter to be adjusted later on. The distance (3) extends in a natural way to a Wasserstein distance between probability measures by (4) d(µ ,µ )= sup φ(x)µ (dx) φ(x)µ (dx) , 1 2 1 2 (cid:12)Z −Z (cid:12) Lipd(φ)≤1(cid:12) (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) where Lip (φ) denotes the Lipschitz constant of φ in the metric d. By the d Monge–Kantorovich duality [44, 49], the right-hand side of (4) is equivalent to (5) d(µ ,µ )= inf d(x,y)µ(dx,dy). 1 2 µ∈C(µ1,µ2)ZZ (Note that the infimum is actually achieved; see [50], Theorem 4.1.) With these notation, one has the following convergence result: Theorem 2.5. Let ( ) be a Markov semigroup over a Banach space t t≥0 P satisfying Assumptions 2 and 3. Then, there exist constants δ>0, α<1 H and T >0 such that (6) d( ∗µ , ∗µ ) αd(µ ,µ ), PT 1 PT 2 ≤ 1 2 for every pair of probability measures µ , µ on . In particular, ( ) 1 2 t t≥0 H P has a unique invariant measure µ and its transition probabilities converge ⋆ exponentially fast to µ . ⋆ Proof. We will prove the general result by first proving it for delta measures, namely, (7) d( ∗δ , ∗δ ) αd(x,y) Pt x Pt y ≤ for all (x,y) . Once this estimate is proven, we can finish the proof ∈H×H of the general case by the following argument. The Monge–Kantorovich duality yields Q (µ ,µ ) so that d(µ ,µ )= 1 2 1 2 ∈C d(x,y)Q(dx,dy). To complete the proof observe that R d( ∗µ , ∗µ ) d( ∗δ , ∗δ )Q(dx,dy) Pt 1 Pt 2 ≤Z Pt x Pt y α d(x,y)Q(dx,dy)=αd(µ ,µ ). 1 2 ≤ Z SPECTRAL GAPS IN WASSERSTEINDISTANCES 9 Let us first show that (7) holds when x y δ for some appropriately k − k≤ chosen δ. Note that by (4) this is equivalent to showing that (8) φ(x) φ(y) αd(x,y)d=efαδ−1 x y t t |P −P |≤ k − k forallsmoothφwithLip (φ) 1.Notethatwecanassumeφ(0)=0,sothat d ≤ this implies that Dφ δ−1 and φ 1. It follows from Assumption 2 ∞ ∞ k k ≤ k k ≤ that D φ C +α δ−1 for every t T . Choosing δ =(1 α )/(2C) t ∞ 1 1 1 k P k ≤ ≥ − and substituting in for C, we get D φ δ−1(1+α )/2, so that (8) t ∞ 1 k P k ≤ holds for t T and α (1+α )/2. 1 1 ≥ ≥ Let us now turn to the case x y >δ. It follows from Assumption 3 k − k that for every t>T (δ) there exists a positive constant a so that for any 2 (x,y) 2 there exists Γ ( ∗δ , ∗δ ) such that Γ(∆ )>a>0, where ∈H ∈C Pt x Pt y δ ∆ = (x′,y′): x′ y′ 1δ . δ { k − k≤ 2 } Since d(x′,y′) 1 on ∆ and d(x′,y′) 1 on the complement, one has ≤ 2 δ ≤ 1 1 a d(x′,y′)Γ(dx′,dy′) Γ(∆ )+1 Γ(∆ )=1 Γ(∆ ) 1 . δ δ δ Z ≤ 2 − − 2 ≤ − 2 Since we are in the setting d(x,y)=1, this implies that when x y >δ, k − k φ(x) φ(y) αd(x,y) t t |P −P |≤ holds for α 1 a and t T (δ). ≥ − 2 ≥ 2 Setting α=max 1 a,1(1+α ) and T =max T ,T (δ) completes the { − 2 2 1 } { 1 2 } proof. (cid:3) Corollary 2.6. Let be as in Theorem 2.5. Then, there exist con- t P stants α<1 and T >0 such that (9) φ µ φ α φ , φ =sup(φ(x) + Dφ(x) ), T ⋆ 1,∞ 1,∞ 1,∞ kP − k ≤ k k k k | | k k x∈H for every Fr´echet differentiable function φ: R. H→ Proof. Defined (x,y)=1 x y .Sincedisequivalenttod ,(6)still 1 1 ∧k − k holds for arbitrary α (but with a different value for T) with d replaced by d .Theclaim thenfollows fromtheMonge–Kantorovich duality, notingthat 1 Lip (φ) 2 φ and, for functions φ with φ(x)µ (dx)=0, φ d1 ≤ k k1,∞ ⋆ k k1,∞ ≤ Lip (φ). (cid:3) R d1 10 M. HAIRERANDJ. MATTINGLY 2.2. A more pathwise perspective. In [20, 37, 38], the authors advocated a pathwise point of view which explicitly constructed coupled versions of the process starting from two different initial conditions in such a way that the two coupled processes converged together exponentially fast. This point of view is very appealing as it conveys a lot of intuition; however, writing down the details can become a bit byzantine. Hence the authors prefer the arguments given in the preceding section for their succinctness and ease of verification. Nonetheless,thecalculations ofthepresentsection providedthe intuition whichguidedtheprevioussection; andhence,wefinditinstructive topresentthem.As noneoftherestof thepaperusesany ofthecalculations fromthissection,wedonotprovideallofthedetails.Ourgoalistoshowhow the estimates from the previous section can be used to construct an explicit coupling in which the expectation of the distance between the trajectories starting from x and y converges to zero exponentially in time. 0 0 Fix a t max(T ,T ), where T and T are the constants in Assumptions 1 2 1 2 ≥ 2and3.Fix δ 0 aswe didin theproofof Theorem2.5.Now for k=0,1,... ≥ define the following sequence of stopping times: r =inf m s :m N, x y (1 α )δ , k k−1 mt mt 1 { ≥ ∈ k − k≤ − } s =inf m r :m N, x y >δ , k k mt mt { ≥ ∈ k − k } where s =0 by definition. −1 If n [r ,s ), let the distribution of (x ,y ) be given by a cou- k k t(n+1) t(n+1) pling w∈hich minimizes the Ed(x ,y ). Hence choosing δ = (1 t(n+1) t(n+1) − α )/(2C) and α ((1+α )/2,1) as in the paragraph below (8), Monge– 1 1 ∈ Kantorovich duality gives that (10) E(d(x ,y )(x ,y )) αδ−1 x y t(n+1) t(n+1) tn tn tn tn | ≤ k − k provided n [r ,s ). Given a random variable X and events A and B, k k for notation∈al convenience, we define E(X;A) = E(X1 ) and P(A;B) = A E(1 1 ), where 1 is the indicator function on the event A. Observe that A B A if x y (1 α )δ, then 0 0 1 k − k≤ − (11) E( x y ;s >n+1(x ,y )) α x y 1 , k t(n+1)− t(n+1)k 1 | tn tn ≤ k tn− tnk s1>n which implies that E( x y ;s >n+1) αn+1 x y . t(n+1) t(n+1) 1 0 0 k − k ≤ k − k From this we see that as long as x and y stay in a δ ball of each other, they will converge toward each other exponentially in expectation. Observe furthermore that δP(s =n)=δP( x y >δ;s >n 1) 1 tn tn 1 k − k − δE(d(x ,y );s >n 1) tn tn 1 ≤ − =δE(E(d(x ,y ) (x ,y ));s >n 1) tn tn t(n−1) t(n−1) 1 | − αE( x y ;s >n 1) αn x y , t(n−1) t(n−1) 1 0 0 ≤ k − k − ≤ k − k

See more

The list of books you might like

Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.