This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Global-in-time mean-field convergence for singular Riesz-type diffusive flows

Matthew Rosenzweig [email protected]  and  Sylvia Serfaty [email protected]
Abstract.

We consider the mean-field limit of systems of particles with singular interactions of the type log|x|-\log|x| or |x|s|x|^{-s}, with 0<s<d20<s<d-2, and with an additive noise in dimensions d3d\geq 3. We use a modulated-energy approach to prove a quantitative convergence rate to the solution of the corresponding limiting PDE. When s>0s>0, the convergence is global in time, and it is the first such result valid for both conservative and gradient flows in a singular setting on d{\mathbb{R}}^{d}. The proof relies on an adaptation of an argument of Carlen-Loss [CL95] to show a decay rate of the solution to the limiting equation, and on an improvement of the modulated-energy method developed in [Due16, Ser20, NRS21], making it so that all prefactors in the time derivative of the modulated energy are controlled by a decaying bound on the limiting solution.

M.R. is supported by the Simons Foundation through the Simons Collaboration on Wave Turbulence and by NSF grant DMS-2052651.
S.S. is supported by NSF grant DMS-2000205 and by the Simons Foundation through the Simons Investigator program.

1. Introduction

1.1. The problem

We consider the first-order mean-field dynamics of stochastic interacting particle systems of the form

(1.1) {dxit=1N1jN:ji𝕄𝗀(xitxjt)dt+2σdWitxit|t=0=xi0i{1,,N}.\begin{cases}dx_{i}^{t}=\displaystyle\frac{1}{N}\sum_{1\leq j\leq N:j\neq i}{\mathbb{M}}\nabla{\mathsf{g}}(x_{i}^{t}-x_{j}^{t})dt+\sqrt{2\sigma}dW_{i}^{t}\\ x_{i}^{t}|_{t=0}=x_{i}^{0}\end{cases}\qquad i\in\{1,\ldots,N\}.

Above, xi0dx_{i}^{0}\in{\mathbb{R}}^{d} are the pairwise distinct initial positions, 𝕄{\mathbb{M}} is a d×dd\times d matrix such that

(1.2) 𝕄ξξ0ξd,{\mathbb{M}}\xi\cdot\xi\leq 0\qquad\forall\xi\in{\mathbb{R}}^{d},

and W1,,WNW_{1},\ldots,W_{N} are independent standard Brownian motions in d{\mathbb{R}}^{d}, so that the noise in (1.1) is of so-called additive type. There are several choices for 𝕄{\mathbb{M}}. For instance, choosing 𝕄=𝕀{\mathbb{M}}=-\mathbb{I} yields gradient-flow/dissipative dynamics, while choosing 𝕄{\mathbb{M}} to be antisymmetric yields Hamiltonian/conservative dynamics. Mixed flows are also permitted. The potential 𝗀{\mathsf{g}} is assumed to be repulsive, which, as we shall later show in Section 4, ensures that there is a unique, global strong solution to the system (1.1). In particular, with probability one, the particles never collide. The model case for 𝗀{\mathsf{g}} is either a logarithmic or Riesz potential indexed by a parameter 0s<d20\leq s<d-2, according to

(1.3) 𝗀(x)={log|x|,s=0|x|s,0<s<d2.{\mathsf{g}}(x)=\begin{cases}-\log|x|,&{s=0}\\ |x|^{-s},&{0<s<d-2}.\end{cases}

The above restriction on ss means that we are considering potentials that are sub-Coulombic: their singularity is below that of the Coulomb potential, which corresponds to s=d2s=d-2. As explained precisely in the next subsection, we can consider a general class of potentials 𝗀{\mathsf{g}} which have sub-Coulombic-type behavior.

Systems of the form (1.1) have numerous applications in the physical and life sciences as well as economics. Examples include vortices in viscous fluids [Ons49, Cho73, Osa87c, MP12], models of the collective motion of microscopic organisms [OS97, TBL06, Per07, GQn15, LY16, FJ17], aggregation phenomena [BGM10, CGM08, Mal03], and opinion dynamics [HK02, Kra00, MT11, XWX11]. For more discussion on applications, we refer the reader to the survey of Jabin and Wang [JW17] and references therein.

Establishing the mean-field limit consists of showing the convergence in a suitable topology as NN\to\infty of the empirical measure

(1.4) μNt1Ni=1Nδxit\mu_{N}^{t}\coloneqq\frac{1}{N}\sum_{i=1}^{N}\delta_{x_{i}^{t}}

associated to a solution x¯Nt(x1t,,xNt)\underline{x}_{N}^{t}\coloneqq(x_{1}^{t},\dots,x_{N}^{t}) of the system (1.1). We remark that for fixed tt, the empirical measure is a random Borel probability measure on d{\mathbb{R}}^{d}. If the points xi0x_{i}^{0}, which themselves depend on NN, are such that μN0\mu_{N}^{0} converges to some regular measure μ0\mu^{0}, then a formal calculation using Itô’s lemma leads to the expectation that for t>0t>0, μNt\mu_{N}^{t} converges to the solution of the Cauchy problem with initial datum μ0\mu^{0} of the limiting evolution equation

(1.5) {tμ=div(μ𝕄𝗀μ)+σΔμμ|t=0=μ0(t,x)+×d\begin{cases}{\partial}_{t}\mu=-\operatorname{\mathrm{div}}(\mu{\mathbb{M}}\nabla{\mathsf{g}}\ast\mu)+\sigma\Delta\mu\\ \mu|_{t=0}=\mu^{0}\end{cases}\qquad(t,x)\in{\mathbb{R}}_{+}\times{\mathbb{R}}^{d}

as the number of particles NN\rightarrow\infty. While the underlying NN-body dynamics are stochastic, we emphasize that the equation (1.5) is completely deterministic, and the noise has been averaged out to become diffusion. Proving the convergence of the empirical measure is closely related to proving propagation of molecular chaos (see [Gol16, HM14, Jab14] and references therein): if fN0(x1,,xN)f_{N}^{0}(x_{1},\dots,x_{N}) is the initial law of the distribution of the NN particles in d{\mathbb{R}}^{d} and if fN0f_{N}^{0} converges to some factorized law (μ0)N(\mu^{0})^{\otimes N}, then the kk-point marginals fN,ktf_{N,k}^{t} converge for all time to (μt)k(\mu^{t})^{\otimes k}.

The mean-field problem for the system (1.1) with σ>0\sigma>0 and interactions which are regular (e.g. globally Lipschitz) has been understood for many years now [McK67, Szn91, M9́6, Mal03] (see also [BGM10, BCnC11, MMW15, Lac21, DT21] for more recent developments still in the regular case). The classical approach consists in comparing the trajectories of the original system (1.1) to those of a cooked-up symmetric particle system coupled to (1.1). Subsequent work has focused on treating the more challenging singular interactions — initially by compactness-type arguments that yield qualitative convergence [Osa86, Osa87b, Osa87c, FHM14, GQn15, LY16, FJ17, LLY19] and later by more quantitative methods that yield an explicit rate for propagation of chaos [Hol16, JW18, BJW19a, BJW20]. To our knowledge, the best results in the literature can quantitatively prove propagation of chaos for singular interactions up to and including the Coulomb case s=d2s=d-2 for conservative dynamics [JW18] and arbitrary 0s<d0\leq s<d in for dissipative dynamics [BJW19a, BJW20]. Unlike the previous aforementioned works which utilize the noise in an essential way, the methods of [JW18, BJW19a] allow for taking vanishing diffusion: σ=σN0\sigma=\sigma_{N}\geq 0, where σN0\sigma_{N}\rightarrow 0 as NN\rightarrow\infty. We also mention that the recent preprint [WZZ21] has gone beyond the mean-field limit and shown the convergence of the fluctuations of the empirical measure to (1.1) to a generalized Ornstein-Uhlenbeck process for singular potentials including the two-dimensional Coulomb case. These state-of-the-art works are limited to the periodic setting.

When there is no noise in the system (1.1) (i.e. σ=0\sigma=0), much more is known mathematically about the mean-field limit thanks to recent advances that are capable of treating the full potential case s<ds<d. Approaches vary, but they all typically involve finding a good metric to measure the distance between the empirical measure and its expected limit and then proving a Gronwall relation for the evolution of this metric. The \infty-Wasserstein metric allowed to treat the sub-Coulombic case s<d2s<d-2 [Hau09, CCH14]. A Wassertein-gradient-flow approach [CFP12, BO19] can also treat the one-dimensional case using the convexity of the Riesz interaction in (and only in) dimension one. The modulated-energy approach of [Due16, Ser20], inspired by the prior work [Ser17], managed to treat the more difficult Coulomb and super-Coulombic case d2s<dd-2\leq s<d for the model potential (1.3). In very recent work by the authors together with Nguyen [NRS21], this modulated-energy approach has been redeveloped to allow to treat the full range s<ds<d and under fairly general assumptions for the potential 𝗀{\mathsf{g}}. In these works, the modulated energy is a Coulomb/Riesz-based metric that can be understood as a renormalization of the negative-order homogeneous Sobolev norm corresponding to the energy space of the equation (1.5). More precisely, it is defined to be

(1.6) FN(x¯N,μ)(d)2𝗀(xy)d(1Ni=1Nδxiμ)2(x,y),F_{N}(\underline{x}_{N},\mu)\coloneqq\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}{\mathsf{g}}(x-y)d\left\lparen\frac{1}{N}\sum_{i=1}^{N}\delta_{x_{i}}-\mu\right\rparen^{\otimes 2}(x,y),

where we remove the infinite self-interaction of each particle by excising the diagonal \triangle.

Contemporaneous to the development of the modulated-energy approach, Jabin and Wang [JW16, JW18] introduced a relative-entropy method capable of treating the mean-field limit of (1.1) when the interaction is moderately singular and which works well with or without noise. The relative-entropy and modulated-energy approaches were recently combined into a modulated free energy method [BJW19b, BJW19a, BJW20] that allows for treating the mean-field limit of (1.1) in the dissipative case, but not the conservative case, set on the torus and under fairly general assumptions on the interaction, impressively allowing even for attractive potentials (e.g. Patlak-Keller-Segel type).

In this article, we show for the first time that the modulated-energy approach of [Due16, Ser20, NRS21] can be extended to treat the mean-field limit of (1.1) in the sub-Coulombic case 0s<d20\leq s<d-2.111Our ability to use the modulated-energy approach in the sub-Coulombic case crucially relies on our recent work [NRS21] with Nguyen, since the previous works [Due16, Ser20] could only treat this way the Coulomb/super-Coulombic case. No incorporation of the entropy, as in the modulated-free-energy approach of [BJW19b, BJW19a, BJW20] is needed. Moreover, the modulated-energy approach is well-suited for exploiting the dissipation of the limiting equation (1.5) to obtain rates of convergence in NN which are uniform over the entire interval [0,)[0,\infty). In other words, mean-field convergence holds globally in time. At the time of completion of this manuscript, this is, to the best of our knowledge, the first instance of such a result for singular potentials. Obtaining a uniform-in-time convergence is important in both theory and practice — for instance, when using a particle system to approximate the limiting equation or its equilibrium states and for quantifying stochastic gradient methods, such as those used in machine learning for other interaction kernels (for instance, see [CB18, MMN18, RVE18]).

Lastly, we mention that previous uses of the modulated energy in the stochastic setting [Ros20b, NRS21] were limited to the case of multiplicative noise, which behaves very differently in the limit as NN\rightarrow\infty. Most notably, the limiting evolution equation is stochastic.

1.2. Formal idea

Let us sketch our main proof in the model case (1.3). For simplicity of exposition, let us also restrict ourselves to the simpler range d4<s<d2d-4<s<d-2. We note that the modulated energy (1.6) is a real-valued continuous stochastic process. Formally by Itô’s lemma (see Section 6 for the rigorous computation), it satisfies the stochastic differential inequality (cf. [Ser20, Lemma 2.1])

(1.7) ddtFN(x¯Nt,μt)(d)2𝗀(xy)(ut(x)ut(y))d(μNtμt)2(x,y)+σ(d)2Δ𝗀(xy)d(μNtμt)2(x,y)+22σNi=1Nd{xit}𝗀(xity)d(μNtμt)(y)W˙it,\frac{d}{dt}F_{N}(\underline{x}_{N}^{t},\mu^{t})\leq\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}\nabla{\mathsf{g}}(x-y)\cdot\left\lparen u^{t}(x)-u^{t}(y)\right\rparen d(\mu_{N}^{t}-\mu^{t})^{\otimes 2}(x,y)\\ +\sigma\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}\Delta{\mathsf{g}}(x-y)d(\mu_{N}^{t}-\mu^{t})^{\otimes 2}(x,y)+\frac{2\sqrt{2\sigma}}{N}\sum_{i=1}^{N}\int_{{\mathbb{R}}^{d}\setminus\{x_{i}^{t}\}}\nabla{\mathsf{g}}(x_{i}^{t}-y)d(\mu_{N}^{t}-\mu^{t})(y)\cdot\dot{W}_{i}^{t},

where we have set ut𝕄𝗀μtu^{t}\coloneqq{\mathbb{M}}\nabla{\mathsf{g}}\ast\mu^{t}. The third term in the right-hand side (formally) has zero expectation and may be ignored for the purposes of this discussion. The first term is the contribution of the drift and also appears in the deterministic case, but the second term is new and due to the nonzero quadratic variation of Brownian motion. Observe that

(1.8) Δ𝗀(xy)=(ds2)|xy|s2.\Delta{\mathsf{g}}(x-y)=-(d-s-2)|x-y|^{-s-2}.

Since 0s<d20\leq s<d-2 by assumption, we see that Δ𝗀\Delta{\mathsf{g}} is superharmonic and equals a constant multiple of 𝗀~-\tilde{{\mathsf{g}}}, where 𝗀~\tilde{{\mathsf{g}}} is the kernel of the Riesz potential operator (Δ)s+2d2(-\Delta)^{\frac{s+2-d}{2}}. We would like to conclude that the second term in the right-hand side of (1.7) is nonpositive by Plancherel’s theorem and therefore may be discarded, but the excision of the diagonal \triangle obstructs this reasoning. Fortunately, prior work of the second author [Ser20, Proposition 3.3] gives the lower bound

(1.9) (d)2𝗀~(xy)d(μNtμt)2(x,y)1N2i=1N𝗀~(ηi)CμtLNi=1Nηids2\begin{split}\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}\tilde{{\mathsf{g}}}(x-y)d\left\lparen\mu_{N}^{t}-\mu^{t}\right\rparen^{\otimes 2}(x,y)\geq-\frac{1}{N^{2}}\sum_{i=1}^{N}\tilde{{\mathsf{g}}}(\eta_{i})-\frac{C\|\mu^{t}\|_{L^{\infty}}}{N}\sum_{i=1}^{N}\eta_{i}^{d-s-2}\end{split}

for all choices of parameters ηi>0\eta_{i}>0. Here, CC is a constant that just depends on s,ds,d. We emphasize that (1.9) is a functional inequality which holds independently of any underlying dynamics. The choice of ηi\eta_{i} that balances the decay in NN between the two terms in the right-hand side of the inequality (1.9) is the typical interparticle distance N1/dN^{-1/d}. Since the LL^{\infty} norm of μt\mu^{t} is a source of decay and we wish to distribute it between terms, we instead choose

(1.10) ηi=(μtLN)1/d1iN,\eta_{i}=(\|\mu^{t}\|_{L^{\infty}}N)^{-1/d}\qquad\forall 1\leq i\leq N,

so that the right-hand side of (1.9) is bounded from below by

(1.11) CσμtLs+2dNdsd,-C\sigma\|\mu^{t}\|_{L^{\infty}}^{\frac{s+2}{d}}N^{-\frac{d-s}{d}},

providing a bound from above for the corresponding term in (1.7). Note that since μt\mu^{t} is time-dependent, our choice for ηi\eta_{i} above depends on time, a trick previously used by the first author [Ros20a, Ros20b] to study the mean-field limit for point vortices with possible multiplicative noise when μt\mu^{t} belongs to a function space which is invariant or critical under the scaling of the equation.

It remains to consider the first term in the right-hand side of (1.7), which, as previously mentioned, also appears in the deterministic case. This expression has the structure of a commutator which has been renormalized through the exclusion of diagonal in order to accommodate the singularity of the Dirac masses. As shown in [Ros20a, NRS21], one can make this commutator intuition rigorous (see Propositions 5.7 and 5.15 below) and, revisiting the estimates there together with some elementary potential analysis, we are able to optimize the dependence in μL\|\mu\|_{L^{\infty}} of the estimate and show the pathwise and pointwise-in-time bound

(1.12) |(d)2𝗀(xy)(ut(x)ut(y))d(μNtμt)2(x,y)|CμtLs+2d(|FN(x¯Nt,μt)|+(1+μtL)Nβ),\left|\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}\nabla{\mathsf{g}}(x-y)\cdot\left\lparen u^{t}(x)-u^{t}(y)\right\rparen d(\mu_{N}^{t}-\mu^{t})^{\otimes 2}(x,y)\right|\\ \leq C\|\mu^{t}\|_{L^{\infty}}^{\frac{s+2}{d}}\left\lparen|F_{N}(\underline{x}_{N}^{t},\mu^{t})|+(1+\|\mu^{t}\|_{L^{\infty}})N^{-\beta}\right\rparen,

where again C,β>0C,\beta>0 are constants depending only s,ds,d.

Now taking expectations of both sides of (1.7), integrating with respect to time, and using Remark 5.2 below to control |FN(x¯Nt,μt)||F_{N}(\underline{x}_{N}^{t},\mu^{t})| by FN(x¯Nt,μt)F_{N}(\underline{x}_{N}^{t},\mu^{t}), we find that

(1.13) 𝔼(|FN(x¯Nt,μt)|)|FN(x¯N0,μ0)|+CμtLsdNβ+C0tμτLs+2d𝔼(|FN(x¯Nτ,μτ)|)𝑑τ+CσNβ0tμτLs+2d𝑑τ.\begin{split}{\mathbb{E}}(|F_{N}(\underline{x}_{N}^{t},\mu^{t})|)&\leq|F_{N}(\underline{x}_{N}^{0},\mu^{0})|+C\|\mu^{t}\|_{L^{\infty}}^{\frac{s}{d}}N^{-\beta}+C\int_{0}^{t}\|\mu^{\tau}\|_{L^{\infty}}^{\frac{s+2}{d}}{\mathbb{E}}(|F_{N}(\underline{x}_{N}^{\tau},\mu^{\tau})|)d\tau\\ &\phantom{=}+C\sigma N^{-\beta}\int_{0}^{t}\|\mu^{\tau}\|_{L^{\infty}}^{\frac{s+2}{d}}d\tau.\end{split}

The structure of the right-hand side of (1.13) allows us to leverage the decay rate of the solution to (1.5). In Proposition 3.8, we show the decay rate

μtLmin{C(σt)d2,μ0L}.\|\mu^{t}\|_{L^{\infty}}\leq\min\{C(\sigma t)^{-\frac{d}{2}},\|\mu^{0}\|_{L^{\infty}}\}.

This is done by revisiting work of Carlen and Loss [CL95] on the optimal decay of nonlinear viscously damped conservation laws, which was essentially restricted to divergence-free vector fields, and adapting it to treat the case of (1.5).

Once this is done, an application of the Gronwall-Bellman lemma to (1.13) yields a uniform-in-time bound if s>0s>0, while if s=0s=0, long-range effects only allow us to obtain an O(tCσ)O(t^{\frac{C}{\sigma}}) growth estimate.

1.3. Assumptions on the potential and main results

We now state the precise assumptions for the class of interaction potentials we consider. This class corresponds to the sub-Coulombic sub-class of the larger class of potentials considered by the authors in collaboration with Nguyen in [NRS21]. In the statement below and throughout this article, the notation 𝟏()\mathbf{1}_{(\cdot)} denotes the indicator function for the condition ()(\cdot).

For d3d\geq 3 and 0s<d20\leq s<d-2, we assume the following:

  1. (i)
    𝗀(x)=𝗀(x){\mathsf{g}}(x)={\mathsf{g}}(-x)
  2. (ii)
    limx0𝗀(x)=\lim_{x\rightarrow 0}{\mathsf{g}}(x)=\infty
  3. (iii)
    r0>0 such thatΔ𝗀0in B(0,r0)\text{$\exists r_{0}>0$ such that}\quad\Delta{\mathsf{g}}\leq 0\quad{\text{in $B(0,r_{0})$}}
  4. (iv)
    k0,|k𝗀(x)|C(1|x|s+k+|log|x||𝟏s=k=0)xd{0}\forall k\geq 0,\qquad|\nabla^{\otimes k}{\mathsf{g}}(x)|\leq C\left\lparen\frac{1}{|x|^{s+k}}+|\log|x||\mathbf{1}_{s=k=0}\right\rparen\quad\forall{x\in{\mathbb{R}}^{d}\setminus\{0\}}
  5. (v)
    |x||𝗀(x)|+|x|2|2𝗀(x)|C𝗀(x)xB(0,r0){0}|x||\nabla{\mathsf{g}}(x)|+|x|^{2}|\nabla^{\otimes 2}{\mathsf{g}}(x)|\leq C{\mathsf{g}}(x)\quad\forall x\in B(0,r_{0}){\setminus\{0\}}
  6. (vi)
    C1|ξ|ds𝗀^(ξ)C2|ξ|dsξd{0}\frac{C_{1}}{|\xi|^{d-s}}\leq\hat{\mathsf{g}}(\xi)\leq\frac{C_{2}}{|\xi|^{d-s}}\quad{\forall\xi\in{\mathbb{R}}^{d}\setminus\{0\}}

    where ^\hat{\cdot} denotes the Fourier transform.

  7. (vii)
    {cs<1 such that𝗀(x)<cs𝗀(y)x,yB(0,r0)with |y|2|x|,s>0c0>0 such that𝗀(x)𝗀(y)c0x,yB(0,r0)with |y|2|x|,s=0\begin{cases}\text{$\exists c_{s}<1$ such that}\quad{\mathsf{g}}(x)<c_{s}{\mathsf{g}}(y)\quad\forall x,y\in B(0,r_{0})\ \text{with $|y|\geq 2|x|$},&{s>0}\\ \text{$\exists c_{0}>0$ such that}\quad{\mathsf{g}}(x)-{\mathsf{g}}(y)\geq c_{0}\quad\forall x,y\in B(0,r_{0})\ \text{with $|y|\geq 2|x|$},&{s=0}\end{cases}
  8. (viii)

    If d4<s<d2d-4<s<d-2, then we also assume that there is an mm\in{\mathbb{N}} and 𝖦:d+m{\mathsf{G}}:{\mathbb{R}}^{d+m}\rightarrow{\mathbb{R}} such that

    Δ𝗀(x)=𝖦(x,0)(x,0)d+m-\Delta{\mathsf{g}}(x)={\mathsf{G}}(x,0)\quad\forall(x,0)\in{\mathbb{R}}^{d+m}
    (1.14) 𝖦(X)=𝖦(X){\mathsf{G}}(X)={\mathsf{G}}(-X)
    (1.15) r0>0 such thatΔ𝖦(X)0in B(0,r0)d+m\text{$\exists r_{0}>0$ such that}\quad\Delta{\mathsf{G}}(X)\leq 0\quad\text{in $B(0,r_{0})\subset{\mathbb{R}}^{d+m}$}
    (1.16) k0,|k𝖦(X)|C|X|s+2+kXB(0,r0)\forall k\geq 0,\quad|\nabla^{\otimes k}{\mathsf{G}}(X)|\leq\frac{C}{|X|^{s+2+k}}\quad\forall X\in B(0,r_{0})
    (1.17) 𝖦^(Ξ)0Ξd+m{0}.\hat{{\mathsf{G}}}(\Xi)\geq 0\quad\forall\Xi\in{\mathbb{R}}^{d+m}\setminus\{0\}.
  9. (ix)

    In the cases s=d2k0s=d-2k\geq 0, for some positive integer kk, we also assume that the (d)(2k+2)({\mathbb{R}}^{d})^{\otimes(2k+2)}-valued kernel

    𝗄(xy)(xy)(2k+1)𝗀(xy)\mathsf{k}(x-y)\coloneqq(x-y)\otimes\nabla^{\otimes(2k+1)}{\mathsf{g}}(x-y)

    is associated to a Calderón-Zygmund operator.333Sufficient and necessary conditions for this Calderón-Zygmund property are explained in [Gra14a, Section 5.4]. The reader may check that this condition is satisfied if 𝗀{\mathsf{g}} is the Riesz potential |x|2kd|x|^{2k-d}.

  10. (x)

    In all cases,

    𝕄:2𝗀(x)0xd{0},{\mathbb{M}}:\nabla^{\otimes 2}{\mathsf{g}}(x)\geq 0\quad\forall x\in{\mathbb{R}}^{d}\setminus\{0\},

    where :: denotes the Frobenius inner product.

We shall say that any potential 𝗀{\mathsf{g}} satisfying assumptions ix is an admissible potential. We refer to [NRS21, Subsection 1.3] for a discussion of the types of potentials permitted under these assumptions. Compared to that work, only assumptions viii and x are new. The former is to ensure that our modulated-energy method can be applied to 𝗀~\tilde{\mathsf{g}} and thus to the diffusion term in (1.7), while the latter is to ensure that the solutions of (1.5) satisfy the temporal decay bounds of the heat equation. Note that x is automatically satisfied if 𝕄{\mathbb{M}} is antisymmetric. Additionally, if 𝕄=𝕀{\mathbb{M}}=-\mathbb{I} so that we consider gradient-flow dynamics, then x amounts to requiring that 𝗀{\mathsf{g}} is globally superharmonic (i.e. r0=r_{0}=\infty in iii). In general, though, we do not require global superharmonicity except where explicitly stated.

We assume that we are given a filtered probability space (Ω,,(t)t0,)(\Omega,\mathcal{F},(\mathcal{F}_{t})_{t\geq 0},\mathbb{P}) on which a countable collection of independent standard dd-dimensional Brownian motions (Wn)n=1(W_{n})_{n=1}^{\infty} are defined. Moreover, (t)t0(\mathcal{F}_{t})_{t\geq 0} is the complete filtration generated by the Brownian motions. All stochastic processes considered in this article are defined on this probability space.

Let x¯N0(d)N\underline{x}_{N}^{0}\in({\mathbb{R}}^{d})^{N} be an NN-tuple of distinct points in d{\mathbb{R}}^{d}. As shown in Proposition 4.5 (more generally Section 4), there exists a unique, global strong solution x¯N\underline{x}_{N} to the Cauchy problem for (1.1). Moreover, with probability one, the particles xitx_{i}^{t} and xjtx_{j}^{t} never collide on the interval [0,)[0,\infty). Let μ0𝒫(d)L\mu^{0}\in\mathcal{P}({\mathbb{R}}^{d})\cap L^{\infty} be a probability measure with LL^{\infty} density with respect to Lebesgue measure. We abuse notation here and throughout the article by using the same symbol to denote both the measure and its density.

As shown in Proposition 3.8 (more generally Section 3), there is a unique, global solution to the Cauchy problem for (1.5) in the class C([0,);𝒫(d)L(d))C([0,\infty);\mathcal{P}({\mathbb{R}}^{d})\cap L^{\infty}({\mathbb{R}}^{d})). In the case of logarithmic interactions (i.e. s=0s=0), we also assume that μ0\mu^{0} satisfies the logarithmic growth condition dlog(1+|x|)𝑑μ0(x)\int_{{\mathbb{R}}^{d}}\log(1+|x|)d\mu^{0}(x), which is propagated locally uniformly by the evolution (see Remark 3.6).

Since x¯N\underline{x}_{N} is stochastic, {FN(x¯Nt,μt)}t0\{F_{N}(\underline{x}_{N}^{t},\mu^{t})\}_{t\geq 0} is a real-valued stochastic process. It is straightforward to check from the non-collision of the particles and Hölder’s inequality that FN(x¯Nt,μt)F_{N}(\underline{x}_{N}^{t},\mu^{t}) is almost surely finite on the interval [0,)[0,\infty) and a continuous process. Our main theorem is a quantitative estimate for the expected magnitude of FN(x¯Nt,μt)F_{N}(\underline{x}_{N}^{t},\mu^{t}).

The first result of this article is the following functional inequality for the expected magnitude of the modulated energy. In the case 0<s<d20<s<d-2, we get a linear growth estimate, while in the case s=0s=0, we have superlinear growth of size O(tσ+Cσ)O(t^{\frac{\sigma+C}{\sigma}}) as tt\rightarrow\infty.

Theorem 1.1.

Let d3d\geq 3, 0s<d20\leq s<d-2, and σ>0\sigma>0. Let x¯N\underline{x}_{N} be a solution to the system (1.1), and let μC([0,);𝒫(d)L(d))\mu\in C([0,\infty);\mathcal{P}({\mathbb{R}}^{d})\cap L^{\infty}({\mathbb{R}}^{d})) be a solution to the PDE (1.5). If s=0s=0, also assume that dlog(1+|x|)𝑑μ0(x)<\int_{{\mathbb{R}}^{d}}\log(1+|x|)d\mu^{0}(x)<\infty. There exists a constant C>0C>0 depending only s,d,σ,μLs,d,\sigma,\|\mu\|_{L^{\infty}}, and the potential 𝗀{\mathsf{g}} through assumptions iix\mathrm{\ref{ass0}-\ref{ass3'}} and an exponent β>0\beta>0 depending only s,ds,d, such that following holds. For all t0t\geq 0 and all NN sufficiently large depending on μ0L\|\mu^{0}\|_{L^{\infty}}, we have that

(1.18) 𝔼(|FN(x¯Nt,μt)|)C(1+t+tσ+Cσ𝟏s=0)(|FN(x¯N0,μ0)|+Nβ).\begin{split}{\mathbb{E}}(|F_{N}(\underline{x}_{N}^{t},\mu^{t})|)\leq C\left\lparen 1+t+t^{\frac{\sigma+C}{\sigma}}\mathbf{1}_{s=0}\right\rparen\left\lparen|F_{N}(\underline{x}_{N}^{0},\mu^{0})|+N^{-\beta}\right\rparen.\end{split}

Assume now that r0=r_{0}=\infty in assumption iii. In other words, the potential 𝗀{\mathsf{g}} is globally superharmonic, as opposed to just in a neighborhood of the origin. The second result of this article is a functional inequality for the expected magnitude of the modulated energy which yields a global bound in the case 0<s<d20<s<d-2. In the case s=0s=0, we have an almost global bound, in the sense that the growth is O(t1σ+)O(t^{{\frac{1}{\sigma}}^{+}}) as tt\rightarrow\infty, which can be arbitrarily small by choosing the diffusion strength σ\sigma arbitrarily large.

Theorem 1.2.

Impose the same assumptions as Theorem 1.1 with the additional condition that r0=r_{0}=\infty. Choose any exponent ds+2<p\frac{d}{s+2}<p\leq\infty. Then there exist constants C,Cp>0C,C_{p}>0 depending on s,d,σ,μLs,d,\sigma,\|\mu\|_{L^{\infty}}, and the potential 𝗀{\mathsf{g}} through assumptions iix\mathrm{\ref{ass0}-\ref{ass3'}} and exponents βp\beta_{p} depending on s,ds,d, such that the following holds. For all t0t\geq 0 and all NN sufficiently large depending on μ0L\|\mu^{0}\|_{L^{\infty}}, we have that

(1.19) 𝔼(|FN(x¯Nt,μt)|)C(1+(t1σlog(1+t))𝟏s=0)(|FN(x¯N0,μ0)|+CpNβp).{\mathbb{E}}(|F_{N}(\underline{x}_{N}^{t},\mu^{t})|)\leq C\left\lparen 1+\left\lparen t^{\frac{1}{\sigma}}\log(1+t)\right\rparen\mathbf{1}_{s=0}\right\rparen\left\lparen|F_{N}(\underline{x}_{N}^{0},\mu^{0})|+C_{p}N^{-\beta_{p}}\right\rparen.

We close this subsection with some remarks on the statements of Theorems 1.1 and 1.2, further extensions, and interesting questions for future work.

Remark 1.3.

An examination of Sections 7.1 and 7.2 will reveal to the interested reader the precise dependence of the constants C,βC,\beta (Cp,βpC_{p},\beta_{p}) on the norms of μ\mu, ss, dd (on pp), and other underlying parameters. We have omitted the explicit dependence and simplified the statements of our final bounds (see (7.10) and (7.15)) in order to make the results more accessible to the reader.

Additionally, we have not attempted to optimize the regularity/integrability assumptions for μ\mu. One can show that in the case s>0s>0, the linear and global bounds of Theorems 1.1 and 1.2, respectively, still hold if we replace the LL^{\infty} assumption with μLp\mu\in L^{p} for finite pp sufficiently large depending on s,ds,d. This, though, comes at the cost of slower decay in NN.

Remark 1.4.

Sufficient conditions for FN(x¯N0,μ0)F_{N}(\underline{x}_{N}^{0},\mu^{0}) to vanish as NN\rightarrow\infty are that the energy of (1.1) converges to the energy of (1.5) and that μN0μ0\mu_{N}^{0}\xrightharpoonup{*}\mu^{0} in the weak-* topology for 𝒫(d)\mathcal{P}({\mathbb{R}}^{d}). See [Due16, Remark 1.2(c)] for more details.

Remark 1.5.

It is well-known [NRS21, Proposition 2.4] that the modulated energy FN(x¯N,μ)F_{N}(\underline{x}_{N},\mu) controls convergence in negative-order Sobolev spaces. Note that since we are restricted to the sub-Coulombic setting, the extension implicit in the cited proposition can be ignored. Consequently, Theorems 1.1 and 1.2 yield a quantitative bound for the expected squared HsH^{s} norm of μNμ\mu_{N}-\mu, for s<d+22s<-\frac{d+2}{2}, of the form

(1.20) 𝔼(μNtμtHs2)Cρ(t)(|FN(x¯N0,μ0)|+Nβ),{\mathbb{E}}\left\lparen\left\|\mu_{N}^{t}-\mu^{t}\right\|_{H^{s}}^{2}\right\rparen\leq C\rho(t)\left\lparen|F_{N}(\underline{x}_{N}^{0},\mu^{0})|+N^{-\beta}\right\rparen,

where ρ(t)\rho(t) is the time factor from either Theorem 1.1 or Theorem 1.2. From this Sobolev convergence and standard arguments (see [HM14, Section 1]), one deduces convergence in law of the empirical measure μN\mu_{N} to μ\mu.

We can also deduce an explicit rate for propagation of chaos for the system (1.1). Indeed, suppose that x¯N0\underline{x}_{N}^{0} are initially distributed in (d)N({\mathbb{R}}^{d})^{N} according to some probability density fN0f_{N}^{0}. Let fNtf_{N}^{t} denote the law of x¯Nt\underline{x}_{N}^{t}, and let fN,ktf_{N,k}^{t} denote the kk-particle marginal of fNtf_{N}^{t}. Then using for instance [RS16, (7.21)], we see that for any symmetric test function φCc((d)k)\varphi\in C_{c}^{\infty}(({\mathbb{R}}^{d})^{k}),

(1.21) |(d)kφd(fN,kt(μt)k)|Cksupx¯k1(d)k1φ(x¯k1,)Hs(d)(d)NμNtμtHs(d)𝑑fN0,\left|\int_{({\mathbb{R}}^{d})^{k}}\varphi d\left\lparen f_{N,k}^{t}-(\mu^{t})^{\otimes k}\right\rparen\right|\leq Ck\sup_{\underline{x}_{k-1}\in({\mathbb{R}}^{d})^{k-1}}\|\varphi(\underline{x}_{k-1},\cdot)\|_{H^{-s}({\mathbb{R}}^{d})}\int_{({\mathbb{R}}^{d})^{N}}\|\mu_{N}^{t}-\mu^{t}\|_{H^{s}({\mathbb{R}}^{d})}df_{N}^{0},

for any s<d+22s<-\frac{d+2}{2}. Combining (1.21) with (1.20) and using duality now yields an explicit rate for propagation of chaos in HsH^{s} norm.

Remark 1.6.

It is an interesting problem to obtain analogues of Theorems 1.1 and 1.2 when a confining potential 𝖵\mathsf{V} is added to the right-hand sides of (1.1), (1.5). In this case, one has a nontrivial equilibrium for the equation (1.5) as tt\rightarrow\infty, which clearly breaks our proof, in particular the Carlen-Loss argument used to prove Proposition 3.8 below. But we still might expect that the difference μNtμt\mu_{N}^{t}-\mu^{t} decays to zero in a suitable topology as tt\rightarrow\infty and that our argument may be salvaged by incorporating the long-time equilibrium for μt\mu^{t}. We plan to investigate this problem in future work. Note that by using so-called similarity variables (see [GW05, Section 1]), one can obtain a local-in-time bound for the modulated energy from Theorems 1.1 and 1.2 in the model interaction case (1.3) and with a quadratic potential. In fact, using our prior work [NRS21] and by appropriately modifying the similarity coordinates (see [SV14, Section 8]), one can also obtain a local-in-time bound for the log or Riesz case 0s<d0\leq s<d in any dimension d1d\geq 1 with a quadratic potential and without noise.

1.4. Comparison with prior results

At the time of completion of this manuscript, our Theorem 1.2 is, to the best of our knowledge, the first time that a quantitative rate of convergence for the mean-field limit of (1.1) has been shown to hold uniformly on the interval [0,)[0,\infty). Additionally, in the case s=0s=0, the rate of convergence as NN\rightarrow\infty is (up to a logN\log N factor) optimal. Existing results in the literature for singular potentials (e.g. [JW18, BJW19a]) have at least an exponential growth in time due to a reliance on a Gronwall-type argument without exploiting the dissipation of the limiting equation. Those works (e.g. [Mal03, CGM08, DEGZ20, DT21]) that do have a uniform convergence result are restricted to regular potentials and require confinement.

Additionally, our theorem is at the level of the empirical measure for the original SDE dynamics for (1.1), as opposed to their associated Liouville/forward Kolmogorov equations for the joint law of the process x¯Nt=(x1t,,xNt)\underline{x}_{N}^{t}=(x_{1}^{t},\ldots,x_{N}^{t}),

(1.22) tfN=i=1Ndivxi(fN1N1ijN𝕄𝗀(xixj))+σi=1NΔxifN.{\partial}_{t}f_{N}=-\sum_{i=1}^{N}\operatorname{\mathrm{div}}_{x_{i}}\left\lparen f_{N}\frac{1}{N}\sum_{1\leq i\neq j\leq N}{\mathbb{M}}\nabla{\mathsf{g}}(x_{i}-x_{j})\right\rparen+\sigma\sum_{i=1}^{N}\Delta_{x_{i}}f_{N}.

Namely, no randomization of the initial data is needed, although as discussed in Remark 1.5, our result implies convergence of this form as well. This stands in contrast to previous work [JW18, BJW19a] whose starting point is the Liouville equation (1.22).

The costs of the strong bounds we obtain with Theorems 1.1 and 1.2 are two-fold. First, we need somewhat stronger assumptions on the potential 𝗀{\mathsf{g}} than in [JW18, BJW19a]–especially the latter work. The reader may find a detailed comparison in [NRS21, Subsection 1.3]. Second, and more importantly, our results are limited to the sub-Coulombic range 0s<d20\leq s<d-2 and dimensions d3d\geq 3. The Coulomb case is barely just out of the reach. Indeed, the diligent reader will note that if 𝗀{\mathsf{g}} is the Coulomb potential, then Δ𝗀=δ\Delta{\mathsf{g}}=-\delta, which is obviously no longer a function. Thus, the argument we described in the previous subsection to bound the second term in the right-hand side of (1.7) no longer applies. The situation is even worse when s>d2s>d-2 as Δ𝗀>0\Delta{\mathsf{g}}>0, meaning what was previously a dissipation term should now cause the modulated energy to grow in time. We mention again that it is an open problem to prove the mean-field limit of (1.1) in the conservative case and when d2<s<dd-2<s<d.

During the final proofreading of the manuscript of this article, Guillin, Le Bris, and Monmarché posted to the arXiv their preprint [GBM21] showing uniform-in-time propagation of chaos in L1L^{1} norm in the periodic setting 𝕋d{\mathbb{T}}^{d} for singular interactions in the range 0sd20\leq s\leq d-2 and d2d\geq 2 with N12N^{-\frac{1}{2}} rate. In particular, they can treat the viscous vortex model corresponding to the d=2d=2 Coulomb case. Their method is very different from ours, as it is based on the relative-entropy method of Jabin-Wang [JW18]. Moreover, they assume random initial data and work at the level of the Liouville equation, and their interaction kernel is assumed to be integrable, have zero distributional divergence, and equal to the divergence of an LL^{\infty} matrix field. Their result is limited to the periodic setting, due to a need for compactness of domain, and to conservative flows. We also note that they impose much stronger regularity assumptions on their solutions to (1.5) than we do.

1.5. Organization of article

In Section 2, we review some estimates for Riesz potential operators and the interaction potential 𝗀{\mathsf{g}} that are used frequently in the paper.

Section 3 is devoted to the study of the limiting equation (1.5), showing that it is globally well-posed and moreover the LpL^{p} norms of solutions satisfy the optimal decay bounds (see Propositions 3.1 and 3.8).

In Section 4, we show that the NN-particle system (1.1) has well-defined dynamics (see Proposition 4.5) in the sense that there exists a unique strong solution and with probability one, the particles never collide. We also introduce in this section a truncation and stopping time procedure that will be used again later in Section 6.

In Section 5, we review properties of the modulated energy and renormalized commutator estimates from the perspective of our recent joint work [NRS21]. We also prove refinements (see Section 5.2) of the results from that work in the case where 𝗀{\mathsf{g}} is globally superharmonic.

Finally, in Sections 6 and 7, we prove our main results, Theorems 1.1 and 1.2. Section 6 gives the rigorous computation of the Itô equation (cf. (1.7)) satisfied by the process FN(x¯Nt,μt)F_{N}(\underline{x}_{N}^{t},\mu^{t}). The main result is Proposition 6.3, which establishes an integral inequality for 𝔼(|FN(x¯Nt,μt)|){\mathbb{E}}(|F_{N}(\underline{x}_{N}^{t},\mu^{t})|). Using this inequality together with the decay bound of Proposition 3.8 and the results of Section 5, we close our Gronwall argument in Section 7, completing the proofs of Theorems 1.1 and 1.2.

1.6. Notation

We close the introduction with the basic notation used throughout the article without further comment.

Given nonnegative quantities AA and BB, we write ABA\lesssim B if there exists a constant C>0C>0, independent of AA and BB, such that ACBA\leq CB. If ABA\lesssim B and BAB\lesssim A, we write ABA\sim B. To emphasize the dependence of the constant CC on some parameter pp, we sometimes write ApBA\lesssim_{p}B or ApBA\sim_{p}B. Also ()+(\cdot)_{+} denotes the positive part of a number.

We denote the natural numbers excluding zero by {\mathbb{N}} and including zero by 0{\mathbb{N}}_{0}. Similarly, we denote the positive real numbers by +{\mathbb{R}}_{+}. Given NN\in{\mathbb{N}} and points x1,N,,xN,Nx_{1,N},\ldots,x_{N,N} in some set XX, we will write x¯N\underline{x}_{N} to denote the NN-tuple (x1,N,,xN,N)(x_{1,N},\ldots,x_{N,N}). Given xnx\in{\mathbb{R}}^{n} and r>0r>0, we denote the ball and sphere centered at xx of radius rr by B(x,r)B(x,r) and B(x,r){\partial}B(x,r), respectively. Given a function ff, we denote the support of ff by suppf\operatorname{supp}f. We use the notation kf\nabla^{\otimes k}f to denote the tensor with components i1ikkf{\partial}_{i_{1}\cdots i_{k}}^{k}f.

We denote the space of Borel probability measures on n{\mathbb{R}}^{n} by 𝒫(n)\mathcal{P}({\mathbb{R}}^{n}). When μ\mu is in fact absolutely continuous with respect to Lebesgue measure, we shall abuse notation by writing μ\mu for both the measure and its density function. We denote the Banach space of complex-valued continuous, bounded functions on n{\mathbb{R}}^{n} by C(n)C({\mathbb{R}}^{n}) equipped with the uniform norm \|\cdot\|_{\infty}. More generally, we denote the Banach space of kk-times continuously differentiable functions with bounded derivatives up to order kk by Ck(n)C^{k}({\mathbb{R}}^{n}) equipped with the natural norm, and we define Ck=1CkC^{\infty}\coloneqq\bigcap_{k=1}^{\infty}C^{k}. We denote the subspace of smooth functions with compact support by Cc(n)C_{c}^{\infty}({\mathbb{R}}^{n}). We denote the Schwartz space of functions by 𝒮(n){\mathcal{S}}({\mathbb{R}}^{n}) and the space of tempered distributions by 𝒮(n){\mathcal{S}}^{\prime}({\mathbb{R}}^{n}).

1.7. Acknowledgments

The second author thanks Eric Vanden-Eijnden for helpful comments.

2. Potential estimates

We review some facts about Riesz potential estimates. For a more thorough discussion, we refer to [Ste70, SM93, Gra14a, Gra14b].

Let d1d\geq 1 For s>ds>-d, we define the Fourier multiplier ||s=(Δ)s/2|\nabla|^{s}=(-\Delta)^{s/2} by

(2.1) ((Δ)s/2f)(||sf^()),f𝒮(d).((-\Delta)^{s/2}f)\coloneqq(|\cdot|^{s}\widehat{f}(\cdot))^{\vee},\qquad f\in{\mathcal{S}}({\mathbb{R}}^{d}).

Since, for s(d,0)s\in(-d,0), the inverse Fourier transform of |ξ|s|\xi|^{s} is the tempered distribution

(2.2) 2sΓ(d+s2)πd2Γ(s2)|x|sd,\frac{2^{s}\Gamma(\frac{d+s}{2})}{\pi^{\frac{d}{2}}\Gamma(-\frac{s}{2})}|x|^{-s-d},

it follows that

(2.3) ((Δ)s/2f)(x)=2sΓ(d+s2)πd2Γ(s2)df(y)|xy|s+d𝑑yxd.((-\Delta)^{s/2}f)(x)=\frac{2^{s}\Gamma(\frac{d+s}{2})}{\pi^{\frac{d}{2}}\Gamma(-\frac{s}{2})}\int_{{\mathbb{R}}^{d}}\frac{f(y)}{|x-y|^{s+d}}dy\qquad\forall x\in{\mathbb{R}}^{d}.

For s(0,d)s\in(0,d), we define the Riesz potential operator of order ss by s(Δ)s/2\mathcal{I}_{s}\coloneqq(-\Delta)^{-s/2} on 𝒮(d){\mathcal{S}}({\mathbb{R}}^{d}).

Remark 2.1.

If 0<s<d0<s<d, we see that the model potential (1.3) corresponds to 𝗀{\mathsf{g}} is a constant times the kernel of ds\mathcal{I}_{d-s}. If s=0s=0, then 𝗀{\mathsf{g}} is a multiple of the inverse Fourier transform of the tempered distribution P.V.|ξ|dcδ0(ξ)\PV|\xi|^{-d}-c\delta_{0}(\xi), for some constant cc. The subtraction of the Dirac mass is to cure the singularity of |ξ|d|\xi|^{-d} near the origin.

s\mathcal{I}_{s} extends to a well-defined operator on any LpL^{p} space, the extension also denoted by s\mathcal{I}_{s} with an abuse of notation, as a consequence of the Hardy-Littlewood-Sobolev (HLS) lemma.

Proposition 2.2.

Let d1d\geq 1, s(0,d)s\in(0,d), and 1<p<q<1<p<q<\infty satisfy the relation

(2.4) 1p1q=sd.\frac{1}{p}-\frac{1}{q}=\frac{s}{d}.

Then for all f𝒮(d)f\in{\mathcal{S}}({\mathbb{R}}^{d}),

(2.5) s(f)Lq\displaystyle\|\mathcal{I}_{s}(f)\|_{L^{q}} fLp,\displaystyle\lesssim\|f\|_{L^{p}},
(2.6) s(f)Ldds,\displaystyle\|\mathcal{I}_{s}(f)\|_{L^{\frac{d}{d-s},\infty}} fL1,\displaystyle\lesssim\|f\|_{L^{1}},

where Lr,L^{r,\infty} denotes the weak-LrL^{r} space. Consequently, s\mathcal{I}_{s} has a unique extension to LpL^{p}, for all 1p<1\leq p<\infty.

The next lemma allows us to control the LL^{\infty} norm of s(f)\mathcal{I}_{s}(f) in terms of the L1L^{1} norm and LpL^{p} norm, for some pp depending on s,ds,d. We omit the proof as it is a straightforward application of Hölder’s inequality.

Lemma 2.3.

Let d1d\geq 1, s(0,d)s\in(0,d), and ds<p\frac{d}{s}<p\leq\infty. Then for all fL1(d)Lp(d)f\in L^{1}({\mathbb{R}}^{d})\cap L^{p}({\mathbb{R}}^{d}), it holds that s(f)C(d)\mathcal{I}_{s}(f)\in C({\mathbb{R}}^{d}) and

(2.7) s(f)LfL11dsd(11p)fLpdsd(11p).\|\mathcal{I}_{s}(f)\|_{L^{\infty}}\lesssim\|f\|_{L^{1}}^{1-\frac{d-s}{d(1-\frac{1}{p})}}\|f\|_{L^{p}}^{\frac{d-s}{d(1-\frac{1}{p})}}.

When the convolution with a Riesz potential is replaced by convolution with the log\log potential, we have an analogue of Lemma 2.3.

Lemma 2.4.

Let d1d\geq 1 and 1<p1<p\leq\infty. For all fLp(d)f\in L^{p}({\mathbb{R}}^{d}) with dlog(1+|x|)|f(x)|𝑑x<\int_{{\mathbb{R}}^{d}}\log(1+|x|)|f(x)|dx<\infty, it holds that log||fCloc(d)\log|\cdot|\ast f\in C_{loc}({\mathbb{R}}^{d}) and

(2.8) |(log||f)(x)|(1+|x|)d(p1)plog(1+|x|)+dlog(1+|y|)|f(y)|dyxd.\left|(\log|\cdot|\ast f)(x)\right|\lesssim(1+|x|)^{\frac{d(p-1)}{p}}\log(1+|x|)+\int_{{\mathbb{R}}^{d}}\log(1+|y|)|f(y)|dy\qquad\forall x\in{\mathbb{R}}^{d}.
Remark 2.5.

If 0s<d0\leq s<d, then for any integer 1k<ds1\leq k<d-s, assumption iv implies that |k𝗀||\nabla^{\otimes k}{\mathsf{g}}| is bounded from above by a constant multiple of the kernel of dsk\mathcal{I}_{d-s-k}. Lemma 2.3 implies that

(2.9) k𝗀fLfL11s+kdfLs+kd,fL1(d)L(d).\|\nabla^{\otimes k}{\mathsf{g}}\ast f\|_{L^{\infty}}\lesssim\|f\|_{L^{1}}^{1-\frac{s+k}{d}}\|f\|_{L^{\infty}}^{\frac{s+k}{d}},\qquad f\in L^{1}({\mathbb{R}}^{d})\cap L^{\infty}({\mathbb{R}}^{d}).

We shall use this estimate frequently in the sequel.

3. The mean-field equation

We start by discussing the well-posedness of the Cauchy problem for and asymptotic decay of solutions to the mean-field PDE (1.5). The latter property is strictly a consequence of the diffusion and is the crucial ingredient to obtain a rate of convergence for the mean-field limit beyond the standard exponential bound given by the Gronwall-Bellman lemma. The results of this section are perhaps mathematical folklore. We present them not for claim for originality but since we could not find them conveniently stated in the literature.

3.1. Local well-posedness

We start by proving local well-posedness for the equation (1.5) for initial data μ0\mu^{0} in the Banach space XL1(d)L(d)X\coloneqq L^{1}({\mathbb{R}}^{d})\cap L^{\infty}({\mathbb{R}}^{d}). That is, we show existence, uniqueness, and continuous dependence on the initial data. The proof proceeds by a contraction mapping argument for the mild formulation of (1.5). In the next subsection, we will upgrade this local well-posedness to global well-posedness through estimates for the temporal decay of the LpL^{p} norms of the solution.

Let us introduce the mild formulation of equations (1.5), on which we base our notion of solution. With etΔe^{t\Delta} denoting the heat flow, we write

(3.1) μt=etσΔμ00te(tκ)σΔdiv(μκ𝕄𝗀μκ)𝑑κ.\mu^{t}=e^{t\sigma\Delta}\mu^{0}-\int_{0}^{t}e^{(t-\kappa)\sigma\Delta}\operatorname{\mathrm{div}}(\mu^{\kappa}{\mathbb{M}}\nabla{\mathsf{g}}\ast\mu^{\kappa})d\kappa.
Proposition 3.1.

Suppose d3d\geq 3 and 0s<d20\leq s<d-2. Let μ0XL1(d)L(d)\mu^{0}\in X\coloneqq L^{1}({\mathbb{R}}^{d})\cap L^{\infty}({\mathbb{R}}^{d}), and let R>0R>0 be such that μ0XR\|\mu^{0}\|_{X}\leq R. Then there exists a unique solution μC([0,T];X)\mu\in C([0,T];X) to equation (3.1) such that TσR2T\sim\sigma R^{-2} and

(3.2) μC([0,T];X)2R.\|\mu\|_{C([0,T];X)}\leq 2R.

Moreover, if μ10X,μ20XR\|\mu_{1}^{0}\|_{X},\|\mu_{2}^{0}\|_{X}\leq R, then there exists TσR2T^{\prime}\sim\sigma R^{-2} such that their associated solutions μ1,μ2\mu_{1},\mu_{2} satisfy the bound

(3.3) μ1μ2C([0,T];X)2μ10μ20X.\|\mu_{1}-\mu_{2}\|_{C([0,T^{\prime}];X)}\leq 2\|\mu_{1}^{0}-\mu_{2}^{0}\|_{X}.
Proof.

Let R>0R>0, let μ0X\mu^{0}\in X with μ0XR\|\mu^{0}\|_{X}\leq R, and let 𝒯\mathcal{T} denote the map

(3.4) μe()σΔμ00()e(κ)σΔdiv(μκ𝕄𝗀μκ)𝑑κ.\mu\mapsto e^{(\cdot)\sigma\Delta}\mu^{0}-\int_{0}^{(\cdot)}e^{(\cdot-\kappa)\sigma\Delta}\operatorname{\mathrm{div}}(\mu^{\kappa}{\mathbb{M}}\nabla{\mathsf{g}}\ast\mu^{\kappa})d\kappa.

We claim that for any appropriate choice of TT, this map is a contraction on the closed ball of radius 2R2R in C([0,T];X)C([0,T];X). Indeed,

𝒯μC([0,T];X)\displaystyle\|\mathcal{T}\mu\|_{C([0,T];X)} μ0X+0Te(κ)σΔdiv(μκ𝕄𝗀μκ)C([0,T];X)𝑑κ\displaystyle\leq\|\mu^{0}\|_{X}+\int_{0}^{T}\|e^{(\cdot-\kappa)\sigma\Delta}\operatorname{\mathrm{div}}(\mu^{\kappa}{\mathbb{M}}\nabla{\mathsf{g}}\ast\mu^{\kappa})\|_{C([0,T];X)}d\kappa
R+C0T(σκ)1/2μ𝕄𝗀μC([0,T];X)𝑑κ\displaystyle\leq R+C\int_{0}^{T}(\sigma\kappa)^{-1/2}\|\mu{\mathbb{M}}\nabla{\mathsf{g}}\ast\mu\|_{C([0,T];X)}d\kappa
(3.5) R+C(T/σ)1/2μ𝕄𝗀μC([0,T];X).\displaystyle\leq R+C(T/\sigma)^{1/2}\|\mu{\mathbb{M}}\nabla{\mathsf{g}}\ast\mu\|_{C([0,T];X)}.

By Hölder’s inequality and Remark 2.5,

(3.6) μt𝕄𝗀μtLpμtLp𝗀μtLμtLpμtL11s+1dμtLs+1d.\|\mu^{t}{\mathbb{M}}\nabla{\mathsf{g}}\ast\mu^{t}\|_{L^{p}}\leq\|\mu^{t}\|_{L^{p}}\|\nabla{\mathsf{g}}\ast\mu^{t}\|_{L^{\infty}}\lesssim\|\mu^{t}\|_{L^{p}}\|\mu^{t}\|_{L^{1}}^{1-\frac{s+1}{d}}\|\mu^{t}\|_{L^{\infty}}^{\frac{s+1}{d}}.

for any exponent 1p1\leq p\leq\infty. Consequently, if μC([0,T];X)2R\|\mu\|_{C([0,T];X)}\leq 2R, then

(3.7) μ𝕄𝗀μC([0,T];X)R2,\|\mu{\mathbb{M}}\nabla{\mathsf{g}}\ast\mu\|_{C([0,T];X)}\lesssim R^{2},

which implies that

(3.8) 𝒯μC([0,T];X)R+C(T/σ)1/2R2,\|\mathcal{T}\mu\|_{C([0,T];X)}\leq R+C(T/\sigma)^{1/2}R^{2},

for some constant CC depending only on s,ds,d and the potential 𝗀{\mathsf{g}} through assumption iv. Similarly, for μ1,μ2B2RC([0,T];X)\mu_{1},\mu_{2}\in B_{2R}\subset C([0,T];X),

(3.9) 𝒯μ1𝒯μ2C([0,T];X)(T/σ)1/2((μ1μ2)𝕄𝗀μ1C([0,T];X)+μ2𝕄𝗀(μ1μ2)C([0,T];X)).\|\mathcal{T}\mu_{1}-\mathcal{T}\mu_{2}\|_{C([0,T];X)}\lesssim(T/\sigma)^{1/2}\Big{(}\|(\mu_{1}-\mu_{2})\nabla{\mathbb{M}}{\mathsf{g}}\ast\mu_{1}\|_{C([0,T];X)}\\ +\|\mu_{2}{\mathbb{M}}\nabla{\mathsf{g}}\ast(\mu_{1}-\mu_{2})\|_{C([0,T];X)}\Big{)}.

Using inequality (3.6), the preceding right-hand side is \lesssim

(3.10) (T/σ)1/2Rμ1μ2C([0,T];X).(T/\sigma)^{1/2}R\|\mu_{1}-\mu_{2}\|_{C([0,T];X)}.

After a little bookkeeping, we see that there is a constant C>0C>0 such that if

(3.11) C(T/σ)1/2R<1,C(T/\sigma)^{1/2}R<1,

then 𝒯\mathcal{T} is indeed a contraction on the closed ball B2RB_{2R}. Consequently, the contraction mapping theorem implies there is a unique solution to equation (3.1) in C([0,T];X)C([0,T];X).

We can also prove Lipschitz-continuous dependence on the initial data. Indeed, let μi0XR\|\mu_{i}^{0}\|_{X}\leq R for i=1,2i=1,2. Then the preceding result tells us there exist unique solutions μi\mu_{i} in C([0,T];X)C([0,T];X) for some TσR2T\sim\sigma R^{-2} and that μiC([0,T];X)R\|\mu_{i}\|_{C([0,T];X)}\lesssim R. Using inequality (3.9), we find that

(3.12) μ1μ2C([0,T];X)μ10μ20X+C(T/σ)1/2Rμ1μ2C([0,T];X).\|\mu_{1}-\mu_{2}\|_{C([0,T];X)}\leq\|\mu_{1}^{0}-\mu_{2}^{0}\|_{X}+C(T/\sigma)^{1/2}R\|\mu_{1}-\mu_{2}\|_{C([0,T];X)}.

Provided that C(T/σ)1/2R<1C(T/\sigma)^{1/2}R<1, we have the bound

(3.13) μ1μ2C([0,T];X)(1C(T/σ)1/2R)1μ10μ20X.\|\mu_{1}-\mu_{2}\|_{C([0,T];X)}\leq(1-C(T/\sigma)^{1/2}R)^{-1}\|\mu_{1}^{0}-\mu_{2}^{0}\|_{X}.

Remark 3.2.

By a Gronwall argument for the energy

(3.14) k=0nd(1+|x|2)m|kμ(x)|2𝑑x\sum_{k=0}^{n}\int_{{\mathbb{R}}^{d}}(1+|x|^{2})^{m}|\nabla^{\otimes k}\mu(x)|^{2}dx

for arbitrarily large integers m,nm,n\in{\mathbb{N}}, it is easy to see that if the initial datum μ0𝒮(d)\mu^{0}\in{\mathcal{S}}({\mathbb{R}}^{d}), then it remains spatially Schwartz on its lifespan. This property combined with the dependence bound (3.13) allows to approximate solutions in C([0,T];X)C([0,T];X) by Schwartz-class solutions.

Remark 3.3.

For a Schwartz-class solution μ\mu, equation (1.5) and the divergence theorem yield

(3.15) ddtdμt𝑑x=d(div(μt𝕄𝗀μt)+σΔμt)𝑑x=0.\frac{d}{dt}\int_{{\mathbb{R}}^{d}}\mu^{t}dx=\int_{{\mathbb{R}}^{d}}\left\lparen-\operatorname{\mathrm{div}}(\mu^{t}{\mathbb{M}}\nabla{\mathsf{g}}\ast\mu^{t})+\sigma\Delta\mu^{t}\right\rparen dx=0.

So by approximation, solutions μC([0,T];X)\mu\in C([0,T];X) obey conservation of mass.

Remark 3.4.

If μC([0,T];X)\mu\in C([0,T];X), then for any 1p1\leq p\leq\infty, it holds that μtLpμtLp\|\mu^{t}\|_{L^{p}}\leq\|\mu^{t^{\prime}}\|_{L^{p}} for all 0ttT0\leq t^{\prime}\leq t\leq T. Indeed, suppose that μ\mu is Schwartz-class and p1p\geq 1 is finite. Then using equation (1.5), we see that

(3.16) ddtμtLpp\displaystyle\frac{d}{dt}\|\mu^{t}\|_{L^{p}}^{p} =pd|μt|p2μtdiv(μt𝕄𝗀μt)𝑑x+pσd|μt|p2μtΔμt𝑑x.\displaystyle=-p\int_{{\mathbb{R}}^{d}}|\mu^{t}|^{p-2}\mu^{t}\operatorname{\mathrm{div}}\left\lparen\mu^{t}{\mathbb{M}}\nabla{\mathsf{g}}\ast\mu^{t}\right\rparen dx+p\sigma\int_{{\mathbb{R}}^{d}}|\mu^{t}|^{p-2}\mu^{t}\Delta\mu^{t}dx.

It follows from integration by parts and the product rule that

(3.17) d|μt|p2μtΔμt𝑑x=d((p2)(|μt|p4μtμt)μt+|μt|p2μt)μtdx0.\int_{{\mathbb{R}}^{d}}|\mu^{t}|^{p-2}\mu^{t}\Delta\mu^{t}dx=-\int_{{\mathbb{R}}^{d}}\left\lparen(p-2)(|\mu^{t}|^{p-4}\mu^{t}\nabla\mu^{t})\mu^{t}+|\mu^{t}|^{p-2}\nabla\mu^{t}\right\rparen\cdot\nabla\mu^{t}dx\leq 0.

Similarly,

(3.18) d|μt|p2μtdiv(μt𝕄𝗀μt)𝑑x=d((p2)(|μt|p4μtμt)μt+|μt|p2μt)μt𝕄𝗀μtdx.-\int_{{\mathbb{R}}^{d}}|\mu^{t}|^{p-2}\mu^{t}\operatorname{\mathrm{div}}\left\lparen\mu^{t}{\mathbb{M}}\nabla{\mathsf{g}}\ast\mu^{t}\right\rparen dx\\ =\int_{{\mathbb{R}}^{d}}\left\lparen(p-2)(|\mu^{t}|^{p-4}\mu^{t}\nabla\mu^{t})\mu^{t}+|\mu^{t}|^{p-2}\nabla\mu^{t}\right\rparen\cdot\mu^{t}{\mathbb{M}}\nabla{\mathsf{g}}\ast\mu^{t}dx.

Writing (|μt|p4μtμt)(μt)2=p1(|μt|p)(|\mu^{t}|^{p-4}\mu^{t}\nabla\mu^{t})(\mu^{t})^{2}=p^{-1}\nabla(|\mu^{t}|^{p}) and |μt|p2μtμt=p1(|μt|p)|\mu^{t}|^{p-2}\mu^{t}\nabla\mu^{t}=p^{-1}\nabla(|\mu^{t}|^{p}), it follows from integrating by parts that

(3.19) pd|μt|p2μtdiv(μt𝕄𝗀μt)𝑑x=(p1)d|μt|pdiv(𝕄𝗀μt)𝑑x0\displaystyle-p\int_{{\mathbb{R}}^{d}}|\mu^{t}|^{p-2}\mu^{t}\operatorname{\mathrm{div}}\left\lparen\mu^{t}{\mathbb{M}}\nabla{\mathsf{g}}\ast\mu^{t}\right\rparen dx=-(p-1)\int_{{\mathbb{R}}^{d}}|\mu^{t}|^{p}\operatorname{\mathrm{div}}({\mathbb{M}}\nabla{\mathsf{g}}\ast\mu^{t})dx\leq 0

where the final inequality follows from assumption x. This takes care of the case p<p<\infty. For p=p=\infty, we take the limit pp\rightarrow\infty^{-}.

Remark 3.5.

Since Remark 3.4 implies the L1L^{1} and LL^{\infty} norms of solutions are nonincreasing and the time of existence in Proposition 3.1 is proportional to μ0X2\|\mu^{0}\|_{X}^{-2}, it follows from iterating Proposition 3.1 that solutions exist globally in C([0,);X)C([0,\infty);X).

Remark 3.6.

Let μ\mu be a nonnegative Schwartz-class solution to (1.5). Then using equation (1.5), integrating by parts, and using the chain rule,

ddtdlog(1+|x|)μt(x)𝑑x\displaystyle\frac{d}{dt}\int_{{\mathbb{R}}^{d}}\log(1+|x|)\mu^{t}(x)dx =dlog(1+|x|)div(μt(𝕄𝗀μt))(x)𝑑x\displaystyle=-\int_{{\mathbb{R}}^{d}}\log(1+|x|)\operatorname{\mathrm{div}}\left\lparen\mu^{t}({\mathbb{M}}\nabla{\mathsf{g}}\ast\mu^{t})\right\rparen(x)dx
+σdlog(1+|x|)Δμt(x)𝑑x\displaystyle\phantom{=}+\sigma\int_{{\mathbb{R}}^{d}}\log(1+|x|)\Delta\mu^{t}(x)dx
=dx|x|(1+|x|)𝕄𝗀μt(x)μt(x)𝑑x\displaystyle=\int_{{\mathbb{R}}^{d}}\frac{x}{|x|(1+|x|)}\cdot{\mathbb{M}}\nabla{\mathsf{g}}\ast\mu^{t}(x)\mu^{t}(x)dx
(3.20) σd(1+|x|)2μt(x)𝑑x.\displaystyle\phantom{=}-\sigma\int_{{\mathbb{R}}^{d}}(1+|x|)^{-2}\mu^{t}(x)dx.

Hence, for any T>0T>0,

dlog(1+|x|)μT(x)𝑑x\displaystyle\int_{{\mathbb{R}}^{d}}\log(1+|x|)\mu^{T}(x)dx dlog(1+|x|)μ0(x)𝑑x+Tsup0tT𝗀μtLμtL1\displaystyle\leq\int_{{\mathbb{R}}^{d}}\log(1+|x|)\mu^{0}(x)dx+T\sup_{0\leq t\leq T}\|\nabla{\mathsf{g}}\ast\mu^{t}\|_{L^{\infty}}\|\mu^{t}\|_{L^{1}}
dlog(1+|x|)μ0(x)𝑑x+Tsup0tTμtL12s+1dμtLs+1d\displaystyle\lesssim\int_{{\mathbb{R}}^{d}}\log(1+|x|)\mu^{0}(x)dx+T\sup_{0\leq t\leq T}\|\mu^{t}\|_{L^{1}}^{2-\frac{s+1}{d}}\|\mu^{t}\|_{L^{\infty}}^{\frac{s+1}{d}}
(3.21) dlog(1+|x|)μ0(x)𝑑x+Tμ0L12s+1dμ0Ls+1d,\displaystyle\leq\int_{{\mathbb{R}}^{d}}\log(1+|x|)\mu^{0}(x)dx+T\|\mu^{0}\|_{L^{1}}^{2-\frac{s+1}{d}}\|\mu^{0}\|_{L^{\infty}}^{\frac{s+1}{d}},

where the penultimate line follows from Remark 2.5 and the ultimate line from the nonincreasing property of LpL^{p} norms. By approximation and continuous dependence, it follows that if μ0X\mu^{0}\in X satisfies dlog(1+|x|)μ0(x)𝑑x<\int_{{\mathbb{R}}^{d}}\log(1+|x|)\mu^{0}(x)dx<\infty, then μt\mu^{t} does as well for all t>0t>0.

Remark 3.7.

Using assumption x, it is not hard to also show that the minimum value of μt\mu^{t} is nondecreasing in time. Consequently, if μ00\mu^{0}\geq 0, then μ0\mu\geq 0 on its lifespan.

3.2. Asymptotic decay

We now show the LpL^{p} norms of the solutions obtained in previous subsection satisfy the same temporal decay estimates as the linear heat equation. This follows the method of [CL95] and extends it to non divergence-free vector fields.

Proposition 3.8.

Suppose that μC([0,);X)\mu\in C([0,\infty);X) is a solution to equation (1.5). If 𝕄:2𝗀0{\mathbb{M}}:\nabla^{\otimes 2}{\mathsf{g}}\neq 0, then assume that μ0\mu\geq 0. Let 1pq1\leq p\leq q\leq\infty. Then for all t>0t>0,

(3.22) μtLq(K(q)K(p))d2(4πσt1/p1/q)d2(1p1q)μ0Lp,\|\mu^{t}\|_{L^{q}}\leq\left\lparen\frac{K(q)}{K(p)}\right\rparen^{\frac{d}{2}}\left\lparen\frac{4\pi\sigma t}{1/p-1/q}\right\rparen^{-\frac{d}{2}(\frac{1}{p}-\frac{1}{q})}\|\mu^{0}\|_{L^{p}},

where

(3.23) K(r)r1/rr1/r,1r.K(r)\coloneqq\frac{{r^{\prime}}^{1/r^{\prime}}}{r^{1/r}},\qquad 1\leq r\leq\infty.

In the conservative case, the velocity field 𝕄𝗀μ{\mathbb{M}}\nabla{\mathsf{g}}\ast\mu is divergence-free and therefore Proposition 3.8 follows from the seminal work [CL95] of Carlen and Loss. In the general case where we only know that the matrix 𝕄{\mathbb{M}} is negative semidefinite (i.e. condition (1.2) holds), we make a small modification of their proof. As it is a crucial ingredient, we first recall the sharp form of Gross’s log-Sobolev inequality [Gro75, CL90, Car91]. Proposition 3.9 below is reproduced from [CL95] (see equation (1.17) in that work).

Proposition 3.9.

Let a>0a>0. Then for all fH1(d)f\in H^{1}({\mathbb{R}}^{d}),

(3.24) d|f(x)|2log(|f(x)|2fL22)𝑑x+(d+dloga2)d|f(x)|2𝑑xaπd|f(x)|2𝑑x.\int_{{\mathbb{R}}^{d}}|f(x)|^{2}\log\left\lparen\frac{|f(x)|^{2}}{\|f\|_{L^{2}}^{2}}\right\rparen dx+\left\lparen d+\frac{d\log a}{2}\right\rparen\int_{{\mathbb{R}}^{d}}|f(x)|^{2}dx\leq\frac{a}{\pi}\int_{{\mathbb{R}}^{d}}|\nabla f(x)|^{2}dx.

Moreover, equality holds if and only if ff is a scalar multiple and translate of fa(x)ad/4eπ|x|2/2af_{a}(x)\coloneqq a^{-d/4}e^{-\pi|x|^{2}/2a}.

Proof of Proposition 3.8.

Using Remark 3.2 and continuous dependence on the initial data, we may assume without loss of generality that μ\mu is spatially Schwartz on its lifespan and μ\mu is CC^{\infty} in time. Therefore, there are no issues of regularity or decay in justifying the computations to follow. Additionally, let us rescale time by defining μσ(t,x)μ(t/σ,x)\mu_{\sigma}(t,x)\coloneqq\mu(t/\sigma,x), which now satisfies the equation

(3.25) tμσ=σ1div(μσ𝕄𝗀μσ)+Δμσ.{\partial}_{t}\mu_{\sigma}=-\sigma^{-1}\operatorname{\mathrm{div}}(\mu_{\sigma}{\mathbb{M}}\nabla{\mathsf{g}}\ast\mu_{\sigma})+\Delta\mu_{\sigma}.

It suffices to show

(3.26) μσtLq(K(q)K(p))d2(4πt1/p1/q)d2(1p1q)μσ0Lp,\|\mu_{\sigma}^{t}\|_{L^{q}}\leq\left\lparen\frac{K(q)}{K(p)}\right\rparen^{\frac{d}{2}}\left\lparen\frac{4\pi t}{1/p-1/q}\right\rparen^{-\frac{d}{2}(\frac{1}{p}-\frac{1}{q})}\|\mu_{\sigma}^{0}\|_{L^{p}},

since replacing tt with σt\sigma t yields the desired result. To simplify the notation, we drop the σ\sigma subscript in what follows and assume that μ\mu solves equation (3.25).

For given p,qp,q as above, let r:[0,T][p,q]r:[0,T]\rightarrow[p,q] be a C1C^{1} increasing function to be specified momentarily. Replacing the absolute value |||\cdot| with (ε2+||2)1/2(\varepsilon^{2}+|\cdot|^{2})^{1/2}, differentiating, then sending ε0+\varepsilon\rightarrow 0^{+}, we find that

(3.27) r(t)2μtLr(t)r(t)1ddtμtLr(t)=r˙(t)d|μt|r(t)log(|μt|r(t)μtLr(t)r(t))𝑑x+r(t)2d|μt|r(t)1sgn(μt)tμtdx.\begin{split}r(t)^{2}\|\mu^{t}\|_{L^{r(t)}}^{r(t)-1}\frac{d}{dt}\|\mu^{t}\|_{L^{r(t)}}&=\dot{r}(t)\int_{{\mathbb{R}}^{d}}|\mu^{t}|^{r(t)}\log\left\lparen\frac{|\mu^{t}|^{r(t)}}{\|\mu^{t}\|_{L^{r(t)}}^{r(t)}}\right\rparen dx\\ &\phantom{=}+r(t)^{2}\int_{{\mathbb{R}}^{d}}|\mu^{t}|^{r(t)-1}\operatorname{sgn}(\mu^{t}){\partial}_{t}\mu^{t}dx.\end{split}

Above, we have used the calculus identity

(3.28) ddtx(t)y(t)=y˙(t)x(t)y(t)logx(t)+y(t)x˙(t)x(t)y(t)1\frac{d}{dt}x(t)^{y(t)}=\dot{y}(t)x(t)^{y(t)}\log x(t)+y(t)\dot{x}(t)x(t)^{y(t)-1}

for C1C^{1} functions x(t)>0x(t)>0 and y(t)y(t). Substituting in equation (1.5), the right-hand side of (3.27) equals

(3.29) r˙(t)d|μt|r(t)log(|μt|r(t)μtLr(t)r(t))𝑑x+r(t)2d|μt|r(t)1sgn(μt)Δμt𝑑xr(t)2σd|μt|r(t)1sgn(μt)div(μt𝕄𝗀μt)𝑑x.\begin{split}&\dot{r}(t)\int_{{\mathbb{R}}^{d}}|\mu^{t}|^{r(t)}\log\left\lparen\frac{|\mu^{t}|^{r(t)}}{\|\mu^{t}\|_{L^{r(t)}}^{r(t)}}\right\rparen dx+r(t)^{2}\int_{{\mathbb{R}}^{d}}|\mu^{t}|^{r(t)-1}\operatorname{sgn}(\mu^{t})\Delta\mu^{t}dx\\ &\phantom{=}-\frac{r(t)^{2}}{\sigma}\int_{{\mathbb{R}}^{d}}|\mu^{t}|^{r(t)-1}\operatorname{sgn}(\mu^{t})\operatorname{\mathrm{div}}(\mu^{t}{\mathbb{M}}\nabla{\mathsf{g}}\ast\mu^{t})dx.\end{split}

By the product rule,

(3.30) r(t)2d|μt|r(t)1sgn(μt)div(μt𝕄𝗀μt)𝑑x=r(t)2d|μt|r(t)div(𝕄𝗀μt)𝑑xr(t)2d|μt|r(t)1sgn(μt)μt(𝕄𝗀μt)𝑑x.-r(t)^{2}\int_{{\mathbb{R}}^{d}}|\mu^{t}|^{r(t)-1}\operatorname{sgn}(\mu^{t})\operatorname{\mathrm{div}}(\mu^{t}{\mathbb{M}}\nabla{\mathsf{g}}\ast\mu^{t})dx\\ =-r(t)^{2}\int_{{\mathbb{R}}^{d}}|\mu^{t}|^{r(t)}\operatorname{\mathrm{div}}({\mathbb{M}}\nabla{\mathsf{g}}\ast\mu^{t})dx-r(t)^{2}\int_{{\mathbb{R}}^{d}}|\mu^{t}|^{r(t)-1}\operatorname{sgn}(\mu^{t})\nabla\mu^{t}\cdot({\mathbb{M}}\nabla{\mathsf{g}}\ast\mu^{t})dx.

We recognize

(3.31) r(t)|μt|r(t)1sgn(μt)μt=(|μt|r(t)).r(t)|\mu^{t}|^{r(t)-1}\operatorname{sgn}(\mu^{t})\nabla\mu^{t}=\nabla(|\mu^{t}|^{r(t)}).

Therefore integrating by parts, the second term in the right-hand side of (3.30) equals

(3.32) r(t)d|μt|r(t)div(𝕄𝗀μt)𝑑x.r(t)\int_{{\mathbb{R}}^{d}}|\mu^{t}|^{r(t)}\operatorname{\mathrm{div}}({\mathbb{M}}\nabla{\mathsf{g}}\ast\mu^{t})dx.

Thus, equality (3.30) and assumption x (and that μ0\mu\geq 0 by assumption if 𝕄:2𝗀{\mathbb{M}}:\nabla^{\otimes 2}{\mathsf{g}} does not vanish on d{0}{\mathbb{R}}^{d}\setminus\{0\}) now give

r(t)2σd|μt|r(t)1sgn(μt)div(μt𝕄𝗀μt)𝑑x\displaystyle-\frac{r(t)^{2}}{\sigma}\int_{{\mathbb{R}}^{d}}|\mu^{t}|^{r(t)-1}\operatorname{sgn}(\mu^{t})\operatorname{\mathrm{div}}(\mu^{t}{\mathbb{M}}\nabla{\mathsf{g}}\ast\mu^{t})dx =r(t)(r(t)1)σd|μt|r(t)(𝕄:2𝗀μt)dx\displaystyle=-\frac{r(t)(r(t)-1)}{\sigma}\int_{{\mathbb{R}}^{d}}|\mu^{t}|^{r(t)}({\mathbb{M}}:\nabla^{\otimes 2}{\mathsf{g}}\ast\mu^{t})dx
(3.33) 0.\displaystyle\leq 0.

Finally, write

(3.34) sgn(μt)=limε0+μtε2+|μt|2.\operatorname{sgn}(\mu^{t})=\lim_{\varepsilon\rightarrow 0^{+}}\frac{\mu^{t}}{\sqrt{\varepsilon^{2}+|\mu^{t}|^{2}}}.

Integrating by parts,

r(t)2d|μt|r(t)1sgn(μt)Δμt𝑑x\displaystyle r(t)^{2}\int_{{\mathbb{R}}^{d}}|\mu^{t}|^{r(t)-1}\operatorname{sgn}(\mu^{t})\Delta\mu^{t}dx
=limε0+(r(t)2(r(t)1)d|μt|r(t)2|μt|ε2+|μt|2|μt|2dx\displaystyle=\lim_{\varepsilon\rightarrow 0^{+}}\Bigg{(}-r(t)^{2}(r(t)-1)\int_{{\mathbb{R}}^{d}}|\mu^{t}|^{r(t)-2}\frac{|\mu^{t}|}{\sqrt{\varepsilon^{2}+|\mu^{t}|^{2}}}|\nabla\mu^{t}|^{2}dx
r(t)2d|μt|r(t)1|μt|2ε2+|μt|2dx+r(t)2d|μt|r(t)1|μt|2|μt|2(ε2+|μt|2)3/2dx)\displaystyle\phantom{=}-r(t)^{2}\int_{{\mathbb{R}}^{d}}|\mu^{t}|^{r(t)-1}\frac{|\nabla\mu^{t}|^{2}}{\sqrt{\varepsilon^{2}+|\mu^{t}|^{2}}}dx+r(t)^{2}\int_{{\mathbb{R}}^{d}}|\mu^{t}|^{r(t)-1}\frac{|\mu^{t}|^{2}|\nabla\mu^{t}|^{2}}{(\varepsilon^{2}+|\mu^{t}|^{2})^{3/2}}dx\Bigg{)}
=r(t)2(r(t)1)d|μt|r(t)2|μt|2𝑑x\displaystyle=-r(t)^{2}(r(t)-1)\int_{{\mathbb{R}}^{d}}|\mu^{t}|^{r(t)-2}|\nabla\mu^{t}|^{2}dx
(3.35) =4(r(t)1)d||μt|r(t)/2|2𝑑x.\displaystyle=-4(r(t)-1)\int_{{\mathbb{R}}^{d}}|\nabla|\mu^{t}|^{r(t)/2}|^{2}dx.

After a little bookkeeping, we realize that we have shown

(3.36) r(t)2μtLr(t)r(t)1ddtμtLr(t)r˙(t)d|μt|r(t)log(|μt|r(t)μtLr(t)r(t))𝑑x4(r(t)1)d||μt|r(t)/2|2𝑑x.\begin{split}r(t)^{2}\|\mu^{t}\|_{L^{r(t)}}^{r(t)-1}\frac{d}{dt}\|\mu^{t}\|_{L^{r(t)}}&\leq\dot{r}(t)\int_{{\mathbb{R}}^{d}}|\mu^{t}|^{r(t)}\log\left\lparen\frac{|\mu^{t}|^{r(t)}}{\|\mu^{t}\|_{L^{r(t)}}^{r(t)}}\right\rparen dx\\ &\phantom{=}-4(r(t)-1)\int_{{\mathbb{R}}^{d}}|\nabla|\mu^{t}|^{r(t)/2}|^{2}dx.\end{split}

The remainder of the proof follows that of Carlen and Loss. We include the details for the sake of completeness.

We apply Proposition 3.9 pointwise in time with choice a=4π(r(t)1)r˙(t)a=\frac{4\pi(r(t)-1)}{\dot{r}(t)} and f=|μt|r(t)/2f=|\mu^{t}|^{r(t)/2} to obtain that

(3.37) r(t)2μtLr(t)r(t)1ddtμtLr(t)r˙(t)(d+d2log(4π(r(t)1)r˙(t)))μtLr(t)r(t).r(t)^{2}\|\mu^{t}\|_{L^{r(t)}}^{r(t)-1}\frac{d}{dt}\|\mu^{t}\|_{L^{r(t)}}\leq-\dot{r}(t)\left\lparen d+\frac{d}{2}\log\left\lparen\frac{4\pi(r(t)-1)}{\dot{r}(t)}\right\rparen\right\rparen\|\mu^{t}\|_{L^{r(t)}}^{r(t)}.

Implicit here is the requirement that r˙(t)>0\dot{r}(t)>0. Define the function

(3.38) G(t)log(μtLr(t)).G(t)\coloneqq\log\left\lparen\|\mu^{t}\|_{L^{r(t)}}\right\rparen.

Then

(3.39) ddtG(t)=1μtLr(t)ddtμtLr(t)r˙(t)r(t)2(d+d2log(4π(r(t)1)r˙(t))).\displaystyle\frac{d}{dt}G(t)=\frac{1}{\|\mu^{t}\|_{L^{r(t)}}}\frac{d}{dt}\|\mu^{t}\|_{L^{r(t)}}\leq-\frac{\dot{r}(t)}{r(t)^{2}}\left\lparen d+\frac{d}{2}\log\left\lparen\frac{4\pi(r(t)-1)}{\dot{r}(t)}\right\rparen\right\rparen.

Set s(t)1/r(t)s(t)\coloneqq 1/r(t), so that the preceding inequality becomes, after writing r1r˙=s(1s)s˙\frac{r-1}{\dot{r}}=-\frac{s(1-s)}{\dot{s}},

(3.40) ddtG(t)s˙(t)(d+d2log(4πs(t)(1s(t))))+d2(s˙(t))log(s˙(t)).\frac{d}{dt}G(t)\leq\dot{s}(t)\left\lparen d+\frac{d}{2}\log\left\lparen 4\pi s(t)(1-s(t))\right\rparen\right\rparen+\frac{d}{2}(-\dot{s}(t))\log(-\dot{s}(t)).

So by the fundamental theorem of calculus,

(3.41) G(t)G(0)0Ts˙(t)(d+d2log(4πs(t)(1s(t))))𝑑td20Ts˙(t)log(s˙(t))𝑑t.\begin{split}G(t)-G(0)&\leq\int_{0}^{T}\dot{s}(t)\left\lparen d+\frac{d}{2}\log\left\lparen 4\pi s(t)(1-s(t))\right\rparen\right\rparen dt\\ &\phantom{=}-\frac{d}{2}\int_{0}^{T}\dot{s}(t)\log\left\lparen-\dot{s}(t)\right\rparen dt.\end{split}

We require that s(0)=1/ps(0)=1/p and s(T)=1/qs(T)=1/q, so by the fundamental theorem of calculus,

(3.42) 0Ts˙(t)(d+d2log(4πs(t)(1s(t))))𝑑t=d2(log(4π)s+log(ss(1s)(1s)))|s=1/ps=1/q.\int_{0}^{T}\dot{s}(t)\left\lparen d+\frac{d}{2}\log(4\pi s(t)(1-s(t)))\right\rparen dt=\frac{d}{2}\left\lparen\log(4\pi)s+\log\left\lparen s^{s}(1-s)^{-(1-s)}\right\rparen\right\rparen|_{s=1/p}^{s=1/q}.

Using the convexity of aalogaa\mapsto a\log a, we minimize the second integral by choosing s(t)s(t) to linearly interpolate between s(0)=1/ps(0)=1/p and s(T)=1/qs(T)=1/q, i.e.

(3.43) s˙(t)=1T(1q1p),0tT.\dot{s}(t)=\frac{1}{T}\left\lparen\frac{1}{q}-\frac{1}{p}\right\rparen,\qquad 0\leq t\leq T.

Thus,

(3.44) d20Ts˙(t)log(s˙(t))𝑑t=d2(1p1q)log(T1/p1/q).-\frac{d}{2}\int_{0}^{T}\dot{s}(t)\log\left\lparen-\dot{s}(t)\right\rparen dt=-\frac{d}{2}\left\lparen\frac{1}{p}-\frac{1}{q}\right\rparen\log\left\lparen\frac{T}{1/p-1/q}\right\rparen.

The desired conclusion now follows from a little bookkeeping and exponentiating both sides of the inequality (3.41). ∎

Remark 3.10.

Our extension of the Carlen-Loss [CL95] method to non-divergence-free vector fields is not limited to proving optimal decay estimates. In fact, it seems we have come across a more general property for which certain parabolic theory is valid. For example, suppose one considers linear equations of the form

(3.45) {tμ=Δμ+div(bμ)+cμμt|t=s=μs(t,x)(s,)×d,\begin{cases}{\partial}_{t}\mu=\Delta\mu+\operatorname{\mathrm{div}}(b\mu)+c\mu\\ \mu^{t}|_{t=s}=\mu^{s}\end{cases}\qquad(t,x)\in(s,\infty)\times{\mathbb{R}}^{d},

where bb is a vector field and cc is a scalar, for simplicity both assumed to be CC^{\infty}. If divb0\operatorname{\mathrm{div}}b\leq 0, then using the same reasoning as in the proof of Proposition 3.8, one can show that the solution μ\mu to (3.45) satisfies the decay estimates

(3.46) μtLq(K(q)K(p))d2(4π(ts)1/p1/q)d2(1p1q)estcκL𝑑κμsLp0<st<.\|\mu^{t}\|_{L^{q}}\leq\left\lparen\frac{K(q)}{K(p)}\right\rparen^{\frac{d}{2}}\left\lparen\frac{4\pi(t-s)}{1/p-1/q}\right\rparen^{-\frac{d}{2}(\frac{1}{p}-\frac{1}{q})}e^{\int_{s}^{t}\|c^{\kappa}\|_{L^{\infty}}d\kappa}\|\mu^{s}\|_{L^{p}}\qquad\forall 0<s\leq t<\infty.

Now one can easily show that equation (3.45) has a (smooth) fundamental solution, and a well-known problem in parabolic theory is to obtain Gaussian upper and lower bounds for such fundamental solutions, since such bounds imply Hölder continuity of the fundamental solution by an argument of Nash [Nas58]. One can adapt the Carlen-Loss argument, which in turn is an adaptation of an earlier argument of Davies [Dav87], to obtain a Gaussian upper bound from (3.46). In certain cases (e.g. [Osa87a, Mae08]), this Gaussian upper bound then implies a corresponding lower bound, and it would seem that these results also hold under the assumption that divb0\operatorname{\mathrm{div}}b\leq 0. We hope to investigate this line of inquiry more in future work.

4. N-particle dynamics

In this section, we discuss the well-posedness of the SDE system (1.1) for fixed NN\in{\mathbb{N}}, in particular that with probability one, the particles never collide. We also discuss stability of the system under regularization of the potential. These regularizations will be important in the sequel when we attempt to apply Itô’s lemma to functions which are singular at the origin.

To prove the well-posedness of the system (1.1), we first consider the well-posedness of the corresponding system where the potential 𝗀{\mathsf{g}} has been smoothly truncated at a short distance ε>0\varepsilon>0 from the origin but otherwise is the same. If the particles remain more than distance ε\varepsilon from one another, then they do not see the truncation and therefore the truncation plays no role. This observation will guide us throughout this subsection.

Let χCc(d)\chi\in C_{c}^{\infty}({\mathbb{R}}^{d}) be a nonnegative bump function satisfying

(4.1) χ(x)={1,|x|1/20,|x|1.\chi(x)=\begin{cases}1,&|x|\leq 1/2\\ 0,&{|x|\geq 1}.\end{cases}

Given ε>0\varepsilon>0, define

(4.2) 𝗀(ε)(x)𝗀(x)(1χ(x/ε)).{\mathsf{g}}_{(\varepsilon)}(x)\coloneqq{\mathsf{g}}(x)(1-\chi(x/\varepsilon)).

The notation 𝗀(ε){\mathsf{g}}_{(\varepsilon)} should not be confused with the notation 𝗀ε{\mathsf{g}}_{\varepsilon} in (5.2) used later in Section 5. By assumption iv, 𝗀(ε)C{\mathsf{g}}_{(\varepsilon)}\in C^{\infty} with

(4.3) 𝗀(ε)L{logε,s=0εs,0<s<d2\|{\mathsf{g}}_{(\varepsilon)}\|_{L^{\infty}}\lesssim\begin{cases}-\log\varepsilon,&{s=0}\\ \varepsilon^{-s},&{0<s<d-2}\end{cases}

and k𝗀(ε)Lε(s+k)\|\nabla^{\otimes k}{\mathsf{g}}_{(\varepsilon)}\|_{L^{\infty}}\lesssim\varepsilon^{-(s+k)} for k1k\geq 1,

(4.4) 𝗀(ε)(x)={𝗀(x),|x|ε0,|x|ε/2.{\mathsf{g}}_{(\varepsilon)}(x)=\begin{cases}{\mathsf{g}}(x),&{|x|\geq\varepsilon}\\ 0,&{|x|\leq\varepsilon/2}.\end{cases}

Define now the truncated version of the system (1.1) by

(4.5) {dxi,ε=1N1jN:ji𝕄𝗀(ε)(xi,εxj,ε)dt+2σdWixi,ε|t=0=xi0.\begin{cases}dx_{i,\varepsilon}=\displaystyle\frac{1}{N}\sum_{1\leq j\leq N:j\neq i}{\mathbb{M}}\nabla{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}-x_{j,\varepsilon})dt+\sqrt{2\sigma}dW_{i}\\ x_{i,\varepsilon}|_{t=0}=x_{i}^{0}.\end{cases}

Since the vector field 𝕄𝗀(ε){\mathbb{M}}\nabla{\mathsf{g}}_{(\varepsilon)} is smooth with bounded derivatives of all order, global well-posedness of (4.5) is trivial. The equality (4.4) implies that if

(4.6) inf0tTmin1ijN|xi,εtxj,εt|ε,\inf_{0\leq t\leq T}\min_{1\leq i\neq j\leq N}|x_{i,\varepsilon}^{t}-x_{j,\varepsilon}^{t}|\geq\varepsilon,

then xi,ε=xix_{i,\varepsilon}=x_{i} on [0,T][0,T] for every 1iN1\leq i\leq N. In other words, the truncated dynamics coincide with the untruncated dynamics, just as remarked at the beginning of the subsection. Accordingly, we can define the stopping time

(4.7) τεinf{0tT:min1ijN|xi,εtxj,εt|2ε},\tau_{\varepsilon}\coloneqq\inf\{0\leq t\leq T:\min_{1\leq i\neq j\leq N}|x_{i,\varepsilon}^{t}-x_{j,\varepsilon}^{t}|\geq 2\varepsilon\},

so that on the random time interval [0,τε(ω)][0,\tau_{\varepsilon}(\omega)], x¯N,ε(ω)x¯N(ω)\underline{x}_{N,\varepsilon}(\omega)\equiv\underline{x}_{N}(\omega), where ωΩ\omega\in\Omega is a sample in the underlying probability space.

Remark 4.1.

For later use, we observe (for instance, see [KS91, Section 3.2.C]) that the quadratic variation of xi,εx_{i,\varepsilon} is the d×dd\times d matrix with components

(4.8) [xi,ε]t,αβ=2σtδαβ,α,β{1,,d},\displaystyle[x_{i,\varepsilon}]^{t,\alpha\beta}=2\sigma t\delta^{\alpha\beta},\qquad\alpha,\beta\in\{1,\ldots,d\},

where δαβ\delta^{\alpha\beta} is the Kronecker δ\delta-function. Similarly, for iji\neq j, the quadratic variation of xi,εxj,εx_{i,\varepsilon}-x_{j,\varepsilon} is given by

(4.9) [xi,εxj,ε]t,αβ=[2σ(WiWj)α,2σ(WiWj)β]t=2σ[Wiα,Wiβ]t+2σ[Wjα,Wjβ]t=4σδαβt.[x_{i,\varepsilon}-x_{j,\varepsilon}]^{t,\alpha\beta}=[\sqrt{2\sigma}(W_{i}-W_{j})^{\alpha},\sqrt{2\sigma}(W_{i}-W_{j})^{\beta}]^{t}=2\sigma[W_{i}^{\alpha},W_{i}^{\beta}]^{t}+2\sigma[W_{j}^{\alpha},W_{j}^{\beta}]^{t}=4\sigma\delta^{\alpha\beta}t.

We first show that with probability one, the particles cannot escape to infinity by controlling the expectation of the moment of inertia

(4.10) IN,εi=1N|xi,ε|2.I_{N,\varepsilon}\coloneqq\sum_{i=1}^{N}|x_{i,\varepsilon}|^{2}.
Lemma 4.2.

There exists a constant C>0C>0 depending only on the dimension dd, such that for all T>0T>0,

(4.11) 𝔼(sup0tTIN,εt)C(IN,ε0+σ(N+T))eCσT.{\mathbb{E}}\left\lparen\sup_{0\leq t\leq T}I_{N,\varepsilon}^{t}\right\rparen\leq C\left\lparen I_{N,\varepsilon}^{0}+\sigma(N+T)\right\rparen e^{C\sigma T}.
Proof.

We sketch the proof. Applying Itô’s lemma with f(x)=|x|2f(x)=|x|^{2}, we find that with probability one,

(4.12) t0,|xi,εt|2|xi0|2=20txi,εκ1jNji𝕄𝗀(xi,εκxj,εκ)dκ+22σ0txi,εκ𝑑Wiκ+2σt.\begin{split}\forall t\geq 0,\qquad|x_{i,\varepsilon}^{t}|^{2}-|x_{i}^{0}|^{2}&=2\int_{0}^{t}x_{i,\varepsilon}^{\kappa}\cdot\sum_{\begin{subarray}{c}1\leq j\leq N\\ j\neq i\end{subarray}}{\mathbb{M}}\nabla{\mathsf{g}}(x_{i,\varepsilon}^{\kappa}-x_{j,\varepsilon}^{\kappa})d\kappa\\ &\phantom{=}+2\sqrt{2\sigma}\int_{0}^{t}x_{i,\varepsilon}^{\kappa}\cdot dW_{i}^{\kappa}+2\sigma t.\end{split}

Since 𝗀\nabla{\mathsf{g}} is odd by assumption i, it follows from the requirement (1.2) that

(4.13) 2i=1Nxi,ε1jNji𝕄𝗀(xi,εxj,ε)=1ijN(xi,εxj,ε)𝕄𝗀(xi,εxj,ε)0.2\sum_{i=1}^{N}x_{i,\varepsilon}\cdot\sum_{\begin{subarray}{c}1\leq j\leq N\\ j\neq i\end{subarray}}{\mathbb{M}}\nabla{\mathsf{g}}(x_{i,\varepsilon}-x_{j,\varepsilon})=\sum_{1\leq i\neq j\leq N}(x_{i,\varepsilon}-x_{j,\varepsilon})\cdot{\mathbb{M}}\nabla{\mathsf{g}}(x_{i,\varepsilon}-x_{j,\varepsilon})\leq 0.

By the Burkholder-Davis-Gundy inequality, denoting again [][\cdot] for the quadratic variation, we have

(4.14) 𝔼(sup0tT|0txi,εκ𝑑Wiκ|)\displaystyle{\mathbb{E}}\left\lparen\sup_{0\leq t\leq T}\left|\int_{0}^{t}x_{i,\varepsilon}^{\kappa}\cdot dW_{i}^{\kappa}\right|\right\rparen 𝔼([0()xi,εκ𝑑Wiκ]T)𝔼(0T|xi,εκ|2𝑑κ).\displaystyle\lesssim{\mathbb{E}}\left\lparen\sqrt{\left[\int_{0}^{(\cdot)}x_{i,\varepsilon}^{\kappa}\cdot dW_{i}^{\kappa}\right]^{T}}\right\rparen\lesssim{\mathbb{E}}\left\lparen\sqrt{\int_{0}^{T}|x_{i,\varepsilon}^{\kappa}|^{2}d\kappa}\right\rparen.

We find after a little bookkeeping that

𝔼(sup0tTIN,εt)\displaystyle{\mathbb{E}}\left\lparen\sup_{0\leq t\leq T}{I}_{N,\varepsilon}^{t}\right\rparen IN,ε0+σ(T+N1/2𝔼(|0TIN,εκ𝑑κ|1/2))\displaystyle\lesssim I_{N,\varepsilon}^{0}+\sigma\left\lparen T+N^{1/2}{\mathbb{E}}\left\lparen\left|\int_{0}^{T}I_{N,\varepsilon}^{\kappa}d\kappa\right|^{1/2}\right\rparen\right\rparen
IN,ε0+σ(T+N1/2𝔼(0TIN,εκ𝑑κ)1/2)\displaystyle\leq I_{N,\varepsilon}^{0}+\sigma\left\lparen T+N^{1/2}{\mathbb{E}}\left\lparen\int_{0}^{T}I_{N,\varepsilon}^{\kappa}d\kappa\right\rparen^{1/2}\right\rparen
(4.15) IN,ε0+σ(T+N+0T𝔼(IN,εκ)𝑑κ),\displaystyle\leq I_{N,\varepsilon}^{0}+\sigma\left\lparen T+N+\int_{0}^{T}{\mathbb{E}}\left\lparen I_{N,\varepsilon}^{\kappa}\right\rparen d\kappa\right\rparen,

where the second line follows from Jensen’s inequality and the third line from the elementary inequality aba2+b22ab\leq\frac{a^{2}+b^{2}}{2} together with Fubini-Tonelli to interchange the expectation with the temporal integration. The desired conclusion now follows from the Gronwall-Bellman lemma. ∎

Remark 4.3.

By Chebyshev’s lemma, Lemma 4.2 implies that with probability one,

(4.16) limRinf{t0:max1ijN|xi,εtxj,εt|R}=.\lim_{R\rightarrow\infty}\inf\{t\geq 0:\max_{1\leq i\neq j\leq N}|x_{i,\varepsilon}^{t}-x_{j,\varepsilon}^{t}|\geq R\}=\infty.

Next, define the function

(4.17) HN,ε(x¯N,ε)1ijN𝗀(ε)(xi,εxj,ε),H_{N,\varepsilon}(\underline{x}_{N,\varepsilon})\coloneqq\sum_{1\leq i\neq j\leq N}{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}-x_{j,\varepsilon}),

which has the interpretation of the energy of the system (4.5).

Lemma 4.4.

There exists a constant C>0C>0 depending only on s,ds,d, and the potential 𝗀{\mathsf{g}} through assumptions iv, v, and vii, such that for all 0<ε<min{12,r02}0<\varepsilon<\min\{\frac{1}{2},\frac{r_{0}}{2}\}, where r0r_{0} is as in iii, and T>0T>0, it holds that

(4.18) (τε<T){(min|x|2ε𝗀(x))1𝔼(CN2(N+eCσT(σ(N+T)+IN,ε0))2+HN,ε(x¯N0)),s=0(min|x|2ε𝗀(x))1𝔼(HN,ε(x¯N0)),0<s<d2.\mathbb{P}(\tau_{\varepsilon}<T)\lesssim\begin{cases}\left\lparen\min_{|x|\leq 2\varepsilon}{\mathsf{g}}(x)\right\rparen^{-1}{\mathbb{E}}\left\lparen\frac{CN^{2}\left\lparen N+e^{C\sigma T}(\sigma(N+T)+I_{N,\varepsilon}^{0})\right\rparen}{2}+H_{N,\varepsilon}(\underline{x}_{N}^{0})\right\rparen,&{s=0}\\ \left\lparen\min_{|x|\leq 2\varepsilon}{\mathsf{g}}(x)\right\rparen^{-1}{\mathbb{E}}(H_{N,\varepsilon}(\underline{x}_{N}^{0})),&{0<s<d-2}.\end{cases}

In particular, by assumption ii, (τε<T)0\mathbb{P}(\tau_{\varepsilon}<T)\rightarrow 0 as ε0+\varepsilon\rightarrow 0^{+}.

Proof.

By Itô’s lemma applied to 𝗀(ε)(xi,εxj,ε){\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}-x_{j,\varepsilon}), it holds with probability one that

(4.19) t0,HN,ε(x¯N,εt)=HN,ε(x¯N,ε0)+21ijN1kNki0t𝕄𝗀(ε)(xi,εκxk,εκ)𝗀(ε)(xi,εκxj,εκ)𝑑κ+2σ1ijN0t𝗀(ε)(xi,εκxj,εκ)d(WiWj)κ+σ1ijN0t2𝗀(ε)(xi,εκxj,εκ):𝕀dκ.\begin{split}\forall t\geq 0,\qquad H_{N,\varepsilon}(\underline{x}_{N,\varepsilon}^{t})&=H_{N,\varepsilon}(\underline{x}_{N,\varepsilon}^{0})\\ &\phantom{=}+2\sum_{1\leq i\neq j\leq N}\sum_{\begin{subarray}{c}1\leq k\leq N\\ k\neq i\end{subarray}}\int_{0}^{t}{\mathbb{M}}\nabla{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}^{\kappa}-x_{k,\varepsilon}^{\kappa})\cdot\nabla{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}^{\kappa}-x_{j,\varepsilon}^{\kappa})d\kappa\\ &\phantom{=}+\sqrt{2\sigma}\sum_{1\leq i\neq j\leq N}\int_{0}^{t}\nabla{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}^{\kappa}-x_{j,\varepsilon}^{\kappa})\cdot d(W_{i}-W_{j})^{\kappa}\\ &\phantom{=}+\sigma\sum_{1\leq i\neq j\leq N}\int_{0}^{t}\nabla^{\otimes 2}{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}^{\kappa}-x_{j,\varepsilon}^{\kappa}):\mathbb{I}\,d\kappa.\end{split}

The second line is nonnegative by condition (1.2) and therefore may be discarded. For the third line, we note that the Itô integral is a martingale with zero initial expectation. So by Doob’s optional sampling theorem,

(4.20) 𝔼(0τε𝗀(ε)(xi,εκxj,εκ)d(WiWj)κ)=0.{\mathbb{E}}\left\lparen\int_{0}^{\tau_{\varepsilon}}\nabla{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}^{\kappa}-x_{j,\varepsilon}^{\kappa})\cdot d(W_{i}-W_{j})^{\kappa}\right\rparen=0.

For the fourth line, we note that

(4.21) 2𝗀(ε)(xi,εκxj,εκ):𝕀=Δ𝗀(ε)(xi,εκxj,εκ).\nabla^{\otimes 2}{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}^{\kappa}-x_{j,\varepsilon}^{\kappa}):\mathbb{I}=\Delta{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}^{\kappa}-x_{j,\varepsilon}^{\kappa}).

Since Δ𝗀0\Delta{\mathsf{g}}\leq 0 by assumption iii and 𝗀=𝗀(ε){\mathsf{g}}={\mathsf{g}}_{(\varepsilon)} outside the ball B(x,ε)B(x,\varepsilon), it follows that

(4.22) Δ𝗀(ε)(x)0|x|ε.\Delta{\mathsf{g}}_{(\varepsilon)}(x)\leq 0\qquad\forall|x|\geq\varepsilon.

Consequently, it holds with probability one that

(4.23) 0τε2𝗀(ε)(xi,εκxj,εκ):𝕀dκ0.\int_{0}^{\tau_{\varepsilon}}\nabla^{\otimes 2}{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}^{\kappa}-x_{j,\varepsilon}^{\kappa}):\mathbb{I}\,d\kappa\leq 0.

Taking expectations of both sides of equation (4.19), we find that

(4.24) 𝔼(HN,ε(x¯N,ετε))𝔼(HN,ε(x¯N0)).{\mathbb{E}}\left\lparen H_{N,\varepsilon}(\underline{x}_{N,\varepsilon}^{\tau_{\varepsilon}})\right\rparen\leq{\mathbb{E}}\left\lparen H_{N,\varepsilon}(\underline{x}_{N}^{0})\right\rparen.

We now want to use this bound to control the minimal distance between particles. To this end, we separately consider the cases s=0s=0 and 0<s<d20<s<d-2. If s=0s=0, we use the moment of inertia to control the possible large negative values of 𝗀{\mathsf{g}}. Using assumptions iv, v, and vii (provided that 2εr02\varepsilon\leq r_{0}), we see that for any t0t\geq 0,

HN,ε(x¯N,εt)\displaystyle H_{N,\varepsilon}(\underline{x}_{N,\varepsilon}^{t}) C1ijN|xi,εtxj,εt|1|log|xi,εtxj,εt||+1ijN|xi,εtxj,εt|2ε𝗀(ε)(xi,εtxj,εt)\displaystyle\geq-C\sum_{\begin{subarray}{c}1\leq i\neq j\leq N\\ |x_{i,\varepsilon}^{t}-x_{j,\varepsilon}^{t}|\geq 1\end{subarray}}\left|\log|x_{i,\varepsilon}^{t}-x_{j,\varepsilon}^{t}|\right|+\sum_{\begin{subarray}{c}1\leq i\neq j\leq N\\ |x_{i,\varepsilon}^{t}-x_{j,\varepsilon}^{t}|\leq 2\varepsilon\end{subarray}}{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}^{t}-x_{j,\varepsilon}^{t})
C1(min|x|2ε𝗀(x))|{(i,j){1,,N}2:ijandε|xi,εtxj,εt|2ε}|\displaystyle\geq C^{-1}\left\lparen\min_{|x|\leq 2\varepsilon}{\mathsf{g}}(x)\right\rparen|\{(i,j)\in\{1,\ldots,N\}^{2}:i\neq j\ \text{and}\ \varepsilon\leq|x_{i,\varepsilon}^{t}-x_{j,\varepsilon}^{t}|\leq 2\varepsilon\}|
(4.25) CN2max(log2,log(i=1N|xi,εt|)).\displaystyle\phantom{=}-CN^{2}\max\left\lparen\log 2,\log(\sum_{i=1}^{N}|x_{i,\varepsilon}^{t}|)\right\rparen.

Note that if i=1N|xi,εt|1\sum_{i=1}^{N}|x_{i,\varepsilon}^{t}|\geq 1, then by Cauchy-Schwarz and since log|x||x|\log|x|\leq|x|,

(4.26) log(i=1N|xi,εt|)N1/2|IN,εt|1/2N+IN,εt2,\log\left\lparen\sum_{i=1}^{N}|x_{i,\varepsilon}^{t}|\right\rparen\leq N^{1/2}|I_{N,\varepsilon}^{t}|^{1/2}\leq\frac{N+I_{N,\varepsilon}^{t}}{2},

where we recall that IN,εI_{N,\varepsilon} is the moment of inertia (4.10). Modulo a null set, this lower bound implies that

{τε<T}\displaystyle\{\tau_{\varepsilon}<T\} {ij{1,,N}2such thatε|xi,ετεxj,ετε|2ε}\displaystyle\subset\{\exists i\neq j\in\{1,\ldots,N\}^{2}\ \text{such that}\ \varepsilon\leq|x_{i,\varepsilon}^{\tau_{\varepsilon}}-x_{j,\varepsilon}^{\tau_{\varepsilon}}|\leq 2\varepsilon\}
(4.27) {HN,ε(x¯N,ετε)CN2(N+IN,ετε)2+C1min|x|2ε𝗀(x)}.\displaystyle\subset\{H_{N,\varepsilon}(\underline{x}_{N,\varepsilon}^{\tau_{\varepsilon}})\geq-\frac{CN^{2}(N+I_{N,\varepsilon}^{\tau_{\varepsilon}})}{2}+C^{-1}\min_{|x|\leq 2\varepsilon}{\mathsf{g}}(x)\}.

So by Chebyshev’s inequality and inequality (4.24),

(4.28) 𝔼(CN2(N+IN,ετε)2)+𝔼(HN,ε(x¯N0))\displaystyle{\mathbb{E}}\left\lparen\frac{CN^{2}(N+I_{N,\varepsilon}^{\tau_{\varepsilon}})}{2}\right\rparen+{\mathbb{E}}\left\lparen H_{N,\varepsilon}(\underline{x}_{N}^{0})\right\rparen C1(min|x|2ε𝗀(x))(τε<T).\displaystyle\geq C^{-1}\left\lparen\min_{|x|\leq 2\varepsilon}{\mathsf{g}}(x)\right\rparen\mathbb{P}(\tau_{\varepsilon}<T).

which in view of Lemma 4.2 concludes the proof if s=0s=0.

If 0<s<d20<s<d-2, then it follows from assumptions iv, v, and vii (provided that 2εr02\varepsilon\leq r_{0})that

HN,ε(x¯N,ε)\displaystyle H_{N,\varepsilon}(\underline{x}_{N,\varepsilon}) 1ijN|xi,εxj,ε|2ε𝗀(ε)(xi,εxj,ε)\displaystyle\geq\sum_{\begin{subarray}{c}1\leq i\neq j\leq N\\ |x_{i,\varepsilon}-x_{j,\varepsilon}|\leq 2\varepsilon\end{subarray}}{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}-x_{j,\varepsilon})
(4.29) CN2+C1(min|x|2ε𝗀(x))|{(i,j){1,,N}2:ijandε|xi,εxj,ε|2ε}|,\displaystyle\geq-CN^{2}+C^{-1}\left\lparen\min_{|x|\leq 2\varepsilon}{\mathsf{g}}(x)\right\rparen|\{(i,j)\in\{1,\ldots,N\}^{2}:i\neq j\ \text{and}\ \varepsilon\leq|x_{i,\varepsilon}-x_{j,\varepsilon}|\leq 2\varepsilon\}|,

which implies that

(4.30) C1(min|x|2ε𝗀(x))(τε<T)𝔼(HN,ε(x¯N0))+CN2,C^{-1}\left\lparen\min_{|x|\leq 2\varepsilon}{\mathsf{g}}(x)\right\rparen\mathbb{P}(\tau_{\varepsilon}<T)\leq{\mathbb{E}}(H_{N,\varepsilon}(\underline{x}_{N}^{0}))+CN^{2},

completing the proof of the lemma. ∎

The next proposition, which is the main result of this section, shows that there is a unique solution to the Cauchy problem for (1.1) in the strong sense.

Proposition 4.5.

With probability one,

(4.31) x¯Ntlimε0+x¯N,εtexistst0,\underline{x}_{N}^{t}\coloneqq\lim_{\varepsilon\rightarrow 0^{+}}\underline{x}_{N,\varepsilon}^{t}\ \text{exists}\ \forall t\geq 0,

and we can unambiguously define x¯N\underline{x}_{N} as the unique strong solution to (1.1). Moreover,

(4.32) (t0,min1ijN|xitxjt|>0)=1.\mathbb{P}\left\lparen\forall t\geq 0,\quad\min_{1\leq i\neq j\leq N}|x_{i}^{t}-x_{j}^{t}|>0\right\rparen=1.
Proof.

Note that if

(4.33) 2εmin1ijN|xi0xj0|,2\varepsilon\leq\min_{1\leq i\neq j\leq N}|x_{i}^{0}-x_{j}^{0}|,

then HN,ε(x¯N0)=HN(x¯N0)H_{N,\varepsilon}(\underline{x}_{N}^{0})=H_{N}(\underline{x}_{N}^{0}). Choose a sequence εk>0\varepsilon_{k}>0 such that

(4.34) k=1(min|x|2εk𝗀(x))1<.\sum_{k=1}^{\infty}\left\lparen\min_{|x|\leq 2\varepsilon_{k}}{\mathsf{g}}(x)\right\rparen^{-1}<\infty.

Then by Lemma 4.4,

(4.35) k=1(τεk<T)<,\sum_{k=1}^{\infty}\mathbb{P}(\tau_{\varepsilon_{k}}<T)<\infty,

so by Borel-Cantelli,

(4.36) (lim supk{τεk<T})=0.\mathbb{P}\left\lparen\limsup_{k\rightarrow\infty}\{\tau_{\varepsilon_{k}}<T\}\right\rparen=0.

Consequently, for almost every sample ωΩ\omega\in\Omega, there exists ε(ω)>0\varepsilon(\omega)>0 such that for all 0<εε(ω)0<\varepsilon\leq\varepsilon(\omega),

(4.37) x¯N,εt(ω)=x¯N,ε(ω)t(ω)on[0,T]andinf0tTmin1ijN|xi,εt(ω)xj,εt(ω)|2ε(ω).\begin{split}\underline{x}_{N,\varepsilon}^{t}(\omega)=\underline{x}_{N,\varepsilon(\omega)}^{t}(\omega)\ \text{on}\ [0,T]\quad\text{and}\quad\inf_{0\leq t\leq T}\min_{1\leq i\neq j\leq N}|x_{i,\varepsilon}^{t}(\omega)-x_{j,\varepsilon}^{t}(\omega)|\geq 2\varepsilon(\omega).\end{split}

Since T>0T>0 was arbitrary, we note that the preceding a.s. statement in fact holds globally in time. ∎

5. Modulated energy and renormalized commutator estimates

In this section, we review the properties of the modulated energy FN(x¯N,μ)F_{N}(\underline{x}_{N},\mu) established in the authors’ joint work with Nguyen [NRS21] along with the associated renormalized commutator estimate proven in that work. These previous results will suffice to prove Theorem 1.1. In the case of potentials which are globally superharmonic (i.e. r0=r_{0}=\infty in assumption iii), we also prove sharper versions (in terms of their μL\|\mu\|_{L^{\infty}} dependence) of the results of [NRS21] that are crucial to obtain global bounds of Theorem 1.2.

Throughout this section, we always assume that μ\mu is a probability measure with density in L(d)L^{\infty}({\mathbb{R}}^{d}). If 0<s<d0<s<d, then it is immediate from Lemma 2.3 that 𝗀μ{\mathsf{g}}\ast\mu is a bounded, continuous function (it is actually Ck,αC^{k,\alpha} for some k0k\in{\mathbb{N}}_{0} and α>0\alpha>0 depending on the value of ss) and therefore the modulated energy is well-defined. If s=0s=0, then we need to impose a suitable decay assumption on μ\mu to compensate for the logarithmic growth of 𝗀{\mathsf{g}} at infinity. For example,

(5.1) dlog(1+|x|)𝑑μ(x)<.\int_{{\mathbb{R}}^{d}}\log(1+|x|)d\mu(x)<\infty.

5.1. Review of results from [NRS21]

We start by reviewing the results of [NRS21, Section 2] on the properties of the modulated energy under the general assumptions on the potential contained in iix. The statements presented below are specialized to the sub-Coulombic setting (i.e. 0sd20\leq s\leq d-2), and the relevant proofs, as well as further comments, may be found in [NRS21, Sections 2, 4].

With r0r_{0} as in iii and 0<η<min{12,r02}0<\eta<\min\{\frac{1}{2},\frac{r_{0}}{2}\}, we let δx(η)\delta_{x}^{(\eta)} denote the uniform probability measure on the sphere B(x,η){\partial}B(x,\eta) and set

(5.2) 𝗀η𝗀δ0(η).{\mathsf{g}}_{\eta}\coloneqq{\mathsf{g}}\ast\delta_{0}^{(\eta)}.

Since 𝗀{\mathsf{g}} is superharmonic in B(0,r0)B(0,r_{0}) by assumption iii, it follows from the formula (d1{\mathcal{H}}^{d-1} is the (d1)(d-1)-dimensional Hausdorff measure in d{\mathbb{R}}^{d})

(5.3) ddrB(x,r)f𝑑d1=1d|B(0,1)|rd1B(x,r)Δf𝑑y,\frac{d}{dr}\mathchoice{{\vbox{\hbox{$\textstyle-$ }}\kern-7.83337pt}}{{\vbox{\hbox{$\scriptstyle-$ }}\kern-5.90005pt}}{{\vbox{\hbox{$\scriptscriptstyle-$ }}\kern-4.75003pt}}{{\vbox{\hbox{$\scriptscriptstyle-$ }}\kern-4.25003pt}}\!\int_{{\partial}B(x,r)}fd{\mathcal{H}}^{d-1}=\frac{1}{d|B(0,1)|r^{d-1}}\int_{B(x,r)}\Delta fdy,

which holds for any sufficiently integrable ff, and an approximation argument that

(5.4) 𝗀η(x)𝗀(x)xB(0,r0η){0}{\mathsf{g}}_{\eta}(x)\leq{\mathsf{g}}(x)\quad\forall x\in B(0,r_{0}-\eta)\setminus\{0\}

and (using assumption iv)

(5.5) |𝗀(x)𝗀η(x)|Cη2|x|s+2|x|2η,|{\mathsf{g}}(x)-{\mathsf{g}}_{\eta}(x)|\leq\frac{C\eta^{2}}{|x|^{s+2}}\qquad\forall|x|\geq 2\eta,

where the constant CC depends on r0r_{0}. By virtue of the mean value inequality (5.4) and assumption iv, the self-interaction of the smeared point mass δx0(η)\delta_{x_{0}}^{(\eta)} satisfies the relation

(5.6) (d)2𝗀(xy)𝑑δx0(η)(x)𝑑δx0(η)(y)=d𝗀η𝑑δ0(η)d𝗀𝑑δ0(η)=𝗀η(0)C(ηs+|logη|𝟏s=0).\int_{({\mathbb{R}}^{d})^{2}}{\mathsf{g}}(x-y)d\delta_{x_{0}}^{(\eta)}(x)d\delta_{x_{0}}^{(\eta)}(y)=\int_{{\mathbb{R}}^{d}}{\mathsf{g}}_{\eta}d\delta_{0}^{(\eta)}\leq\int_{{\mathbb{R}}^{d}}{\mathsf{g}}d\delta_{0}^{(\eta)}={\mathsf{g}}_{\eta}(0)\leq C\left\lparen\eta^{-s}+|\log\eta|\mathbf{1}_{s=0}\right\rparen.

The next result we recall [NRS21, Proposition 2.1] expresses the crucial monotonicity property of the modulated energy with respect to the smearing radii when expressed as the limit

(5.7) FN(x¯N,μ)=limαi0((d)2𝗀(xy)d(1Ni=1Nδxi(αi)μ)2(x,y)1N2i=1Nd𝗀αi𝑑δ0(αi)).F_{N}(\underline{x}_{N},\mu)=\lim_{\alpha_{i}\rightarrow 0}\Bigg{(}\int_{({\mathbb{R}}^{d})^{2}}{\mathsf{g}}(x-y)d\left\lparen\frac{1}{N}\sum_{i=1}^{N}\delta_{x_{i}}^{(\alpha_{i})}-\mu\right\rparen^{\otimes 2}(x,y)-\frac{1}{N^{2}}\sum_{i=1}^{N}\int_{{\mathbb{R}}^{d}}{\mathsf{g}}_{\alpha_{i}}d\delta_{0}^{(\alpha_{i})}\Bigg{)}.

It also shows that the modulated energy is bounded from below, coercive, and controls the microscale interactions [NRS21, Corollary 2.3].

Proposition 5.1.

Let d3d\geq 3 and 0sd20\leq s\leq d-2. Suppose that x¯N(d)N\underline{x}_{N}\in({\mathbb{R}}^{d})^{N} is pairwise distinct and μ𝒫(d)L(d)\mu\in\mathcal{P}({\mathbb{R}}^{d})\cap L^{\infty}({\mathbb{R}}^{d}). In the case s=0s=0, also suppose that dlog(1+|x|)𝑑μ(x)<\int_{{\mathbb{R}}^{d}}\log(1+|x|)d\mu(x)<\infty. There exists a constant CC depending only on s,ds,d and the potential 𝗀{\mathsf{g}} through assumptions iii,iv,vi\mathrm{\ref{ass1},\ref{ass2},\ref{ass3}}, such that for every choice of 0<η1,,ηN<min{12,r02}0<\eta_{1},\ldots,\eta_{N}<\min\{\frac{1}{2},\frac{r_{0}}{2}\},

(5.8) 1N21ijN|xixj|r02(𝗀(xjxi)𝗀ηi(xjxi))++C11Ni=1Nδxi(ηi)μH˙sd22FN(x¯N,μ)+CNi=1N((ηi2+ηis(1+|logηi|𝟏s=0)N)+CμLηids(1+|logηi|(𝟏s=0+𝟏s=d2))).\frac{1}{N^{2}}\sum_{\begin{subarray}{c}1\leq i\neq j\leq N\\ |x_{i}-x_{j}|\leq\frac{r_{0}}{2}\end{subarray}}\left\lparen{\mathsf{g}}(x_{j}-x_{i})-{\mathsf{g}}_{\eta_{i}}(x_{j}-x_{i})\right\rparen_{+}+C^{-1}\left\|\frac{1}{N}\sum_{i=1}^{N}\delta_{x_{i}}^{(\eta_{i})}-\mu\right\|_{\dot{H}^{\frac{s-d}{2}}}^{2}\leq F_{N}(\underline{x}_{N},\mu)\\ +\frac{C}{N}\sum_{i=1}^{N}\Bigg{(}\left\lparen\eta_{i}^{2}+\frac{\eta_{i}^{-s}(1+|\log\eta_{i}|\mathbf{1}_{s=0})}{N}\right\rparen+C\|\mu\|_{L^{\infty}}\eta_{i}^{d-s}\left\lparen 1+|\log\eta_{i}|(\mathbf{1}_{s=0}+\mathbf{1}_{s=d-2})\right\rparen\Bigg{)}.
Remark 5.2.

Since 0sd20\leq s\leq d-2 by assumption, we can balance the error terms in (5.8) by setting

(5.9) ηi2=ηisNηi=N1s+2,\eta_{i}^{2}=\frac{\eta_{i}^{-s}}{N}\Longleftrightarrow\eta_{i}=N^{-\frac{1}{s+2}},

which, in particular, yields the lower bound

(5.10) FN(x¯N,μ)C(1+μL)N2s+2(1+(logN)(𝟏s=0+𝟏s=d2)).F_{N}(\underline{x}_{N},\mu)\geq-C(1+\|\mu\|_{L^{\infty}})N^{-\frac{2}{s+2}}\left\lparen 1+(\log N)(\mathbf{1}_{s=0}+\mathbf{1}_{s=d-2})\right\rparen.
Remark 5.3.

If instead of vi, we only assume that 𝗀^0\hat{{\mathsf{g}}}\geq 0 on d{0}{\mathbb{R}}^{d}\setminus\{0\}, then the proof of [NRS21, Proposition 2.1] yields the bound

(5.11) 1N21ijN|xixj|r02(𝗀(xjxi)𝗀ηi(xjxi))+FN(x¯N,μ)+CNi=1N((ηi2+ηis(1+|logηi|𝟏s=0)N)+CμLηids(1+|logηi|(𝟏s=0+𝟏s=d2))).\frac{1}{N^{2}}\sum_{\begin{subarray}{c}1\leq i\neq j\leq N\\ |x_{i}-x_{j}|\leq\frac{r_{0}}{2}\end{subarray}}\left\lparen{\mathsf{g}}(x_{j}-x_{i})-{\mathsf{g}}_{\eta_{i}}(x_{j}-x_{i})\right\rparen_{+}\leq F_{N}(\underline{x}_{N},\mu)\\ +\frac{C}{N}\sum_{i=1}^{N}\Bigg{(}\left\lparen\eta_{i}^{2}+\frac{\eta_{i}^{-s}(1+|\log\eta_{i}|\mathbf{1}_{s=0})}{N}\right\rparen+C\|\mu\|_{L^{\infty}}\eta_{i}^{d-s}\left\lparen 1+|\log\eta_{i}|(\mathbf{1}_{s=0}+\mathbf{1}_{s=d-2})\right\rparen\Bigg{)}.

The next result we recall [NRS21, Proposition 2.2] concerns the analogue of Proposition 5.1 in the case d2<s<dd-2<s<d. One of the key new insights from [NRS21] is that although superharmonicity may fail in the space d{\mathbb{R}}^{d}, as it does for the Riesz potential |x|s|x|^{-s}, superharmonicity may be restored by considering the potential as the restriction of a potential 𝖦{\mathsf{G}} (i.e. 𝗀(x)=𝖦(x,0){\mathsf{g}}(x)={\mathsf{G}}(x,0)) in an extended space d+m{\mathbb{R}}^{d+m}, where the size of mm depends on the value of ss so as to make 𝖦{\mathsf{G}} superharmonic in a neighborhood of the origin. Namely, suppose that 𝗀:d{0}{\mathsf{g}}:{\mathbb{R}}^{d}\setminus\{0\}\rightarrow{\mathbb{R}} is such that there exists 𝖦:d+m{0}{\mathsf{G}}:{\mathbb{R}}^{d+m}\setminus\{0\}\rightarrow{\mathbb{R}} with 𝗀(x)=𝖦(x,0){\mathsf{g}}(x)={\mathsf{G}}(x,0) and satisfying conditions (1.14) – (1.17). With the notation X=(x,z)d+mX=(x,z)\in{\mathbb{R}}^{d+m} and Xi=(xi,0)X_{i}=(x_{i},0), we let δX(η)\delta_{X}^{(\eta)} denote the uniform probability measure on the sphere B(X,r)d+m{\partial}B(X,r)\subset{\mathbb{R}}^{d+m} and set

(5.12) 𝖦η𝖦δ0(η){\mathsf{G}}_{\eta}\coloneqq{\mathsf{G}}\ast\delta_{0}^{(\eta)}

for 0<η<min{12,r02}0<\eta<\min\{\frac{1}{2},\frac{r_{0}}{2}\}. Analogously to (5.4), (5.5), (5.6), we have

(5.13) 𝖦η(X)𝖦(X)XB(0,r0η){0},{\mathsf{G}}_{\eta}(X)\leq{\mathsf{G}}(X)\qquad\forall X\in B(0,r_{0}-\eta)\setminus\{0\},
(5.14) |𝖦(X)𝖦η(X)|Cη2|X|s+2|X|2η,|{\mathsf{G}}(X)-{\mathsf{G}}_{\eta}(X)|\leq\frac{C\eta^{2}}{|X|^{s+2}}\qquad\forall|X|\geq 2\eta,

and

(5.15) (d+m)2𝖦(xy)𝑑δ0(η)(x)𝑑δ0(η)(y)𝖦η(0)C(ηs+|logη|𝟏d=1s=0).\int_{({\mathbb{R}}^{d+m})^{2}}{\mathsf{G}}(x-y)d\delta_{0}^{(\eta)}(x)d\delta_{0}^{(\eta)}(y)\leq{\mathsf{G}}_{\eta}(0)\leq C\left\lparen\eta^{-s}+|\log\eta|\mathbf{1}_{d=1\wedge s=0}\right\rparen.

Again, the constant CC in (5.14) depends on r0r_{0}.

Proposition 5.4.

Let d3d\geq 3 and d2<s<dd-2<s<d. Let 𝗀,𝖦{\mathsf{g}},{\mathsf{G}} be as above. Suppose that x¯N(d)N\underline{x}_{N}\in({\mathbb{R}}^{d})^{N} is a pairwise distinct configuration and μ𝒫(d)L(d)\mu\in\mathcal{P}({\mathbb{R}}^{d})\cap L^{\infty}({\mathbb{R}}^{d}). There exists a constant C>0C>0 only depending on s,ds,d and on 𝗀,𝖦{\mathsf{g}},{\mathsf{G}}, such that for every 0<η1,,ηN<min{12,r02}0<\eta_{1},\ldots,\eta_{N}<\min\{\frac{1}{2},\frac{r_{0}}{2}\}, we have

(5.16) 1N21ijN|xixj|r02(𝗀(xjxi)𝖦ηi(xjxi,0))+FN(x¯N,μ)+i=1Nηis(1+|logηi|𝟏s=0)N2+CNi=1N(μLηids+ηi2).\frac{1}{N^{2}}\sum_{\begin{subarray}{c}1\leq i\neq j\leq N\\ |x_{i}-x_{j}|\leq\frac{r_{0}}{2}\end{subarray}}\left\lparen{\mathsf{g}}(x_{j}-x_{i})-{\mathsf{G}}_{\eta_{i}}(x_{j}-x_{i},0)\right\rparen_{+}\leq F_{N}(\underline{x}_{N},\mu)+\sum_{i=1}^{N}\frac{\eta_{i}^{-s}(1+|\log\eta_{i}|\mathbf{1}_{s=0})}{N^{2}}\\ +\frac{C}{N}\sum_{i=1}^{N}\left\lparen\|\mu\|_{L^{\infty}}\eta_{i}^{d-s}+\eta_{i}^{2}\right\rparen.
Remark 5.5.

Strictly speaking, the inequality (5.16) differs from [NRS21, (2.23), Proposition 2.2] by the omission of a positive term (a suitable squared norm of μNtμt\mu_{N}^{t}-\mu^{t}). The reason we have omitted this term is because we no longer assume that 𝖦^(Ξ)|Ξ|sdm\hat{{\mathsf{G}}}(\Xi)\sim|\Xi|^{s-d-m}. Instead, condition (1.17) only tells us that

(5.17) (d+m)2𝖦(XY)d(1Ni=1NδXi(ηi)μ~)(X,Y)0,\int_{({\mathbb{R}}^{d+m})^{2}}{\mathsf{G}}(X-Y)d\left\lparen\frac{1}{N}\sum_{i=1}^{N}\delta_{X_{i}}^{(\eta_{i})}-\tilde{\mu}\right\rparen(X,Y)\geq 0,

which is good enough for the purposes of this article.

As an application of Proposition 5.1 if 0sd40\leq s\leq d-4 and Proposition 5.4 if d4<s<d2d-4<s<d-2 using assumption viii, expressions like the second term appearing in the right-hand side of (1.7), which is due to the nonzero quadratic variation of the Brownian motion when we calculate the Itô equation for the modulated energy FN(x¯Nt,μt)F_{N}(\underline{x}_{N}^{t},\mu^{t}), are nonpositive up to a controllable error.

Corollary 5.6.

Let d3d\geq 3 and 0s<d20\leq s<d-2. Suppose that x¯N(d)N\underline{x}_{N}\in({\mathbb{R}}^{d})^{N} is pairwise distinct and μ𝒫(d)L(d)\mu\in\mathcal{P}({\mathbb{R}}^{d})\cap L^{\infty}({\mathbb{R}}^{d}). Let 𝗀{\mathsf{g}} be a potential satisfying assumptions i,iii,iv,vi,viii\mathrm{\ref{ass0},\ref{ass1},\ref{ass2},\ref{ass3},\ref{ass4}}. There exists a constant C>0C>0 depending only on s,ds,d and 𝗀{\mathsf{g}}, such that

(5.18) (d)2(Δ𝗀)(xy)d(1Ni=1Nδxiμ)2(x,y)C(1+μL)Nmin{2,ds2}min{s+4,d}.\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}(-\Delta{\mathsf{g}})(x-y)d\left\lparen\frac{1}{N}\sum_{i=1}^{N}\delta_{x_{i}}-\mu\right\rparen^{\otimes 2}(x,y)\geq-C(1+\|\mu\|_{L^{\infty}})N^{-\frac{\min\{2,d-s-2\}}{\min\{s+4,d\}}}.
Proof.

If 0sd40\leq s\leq d-4, we use (5.11) applied with potential Δ𝗀-\Delta{\mathsf{g}} and with each ηi=N1s+4\eta_{i}=N^{-\frac{1}{s+4}}. If d4<s<d2d-4<s<d-2, we use (5.16) applied with extended potential 𝖦{\mathsf{G}} for Δ𝗀-\Delta{\mathsf{g}} given by assumption viii and with each ηi=N1d\eta_{i}=N^{-\frac{1}{d}}. ∎

Finally, we close this section by recalling [NRS21, Proposition 4.1] the renormalized commutator estimate from that work. As commented in the introduction, such estimates are the main workhorse to close Gronwall arguments based on the modulated energy.

Proposition 5.7.

Let d3d\geq 3 and 0sd20\leq s\leq d-2. Let x¯N(d)N\underline{x}_{N}\in({\mathbb{R}}^{d})^{N} be a pairwise distinct configuration, and μ𝒫(d)L(d)\mu\in\mathcal{P}({\mathbb{R}}^{d})\cap L^{\infty}({\mathbb{R}}^{d}). If s=0s=0, assume that dlog(1+|x|)𝑑μ(x)<\int_{{\mathbb{R}}^{d}}\log(1+|x|)d\mu(x)<\infty. Let vv be a continuous vector field on d{\mathbb{R}}^{d}. There exists a constant CC depending only d,sd,s and on the potential 𝗀{\mathsf{g}} through assumptions ivii, ix, such that

(5.19) |(d)2(v(x)v(y))𝗀(xy)d(1Ni=1Nδxiμ)2(x,y)|C(vL+||ds2vL2dd2s𝟏s<d2)(FN(x¯N,μ)+CNs+3(s+2)(s+1)||s+1dμL+C(1+μL)N2(s+2)(s+1)(1+(logN)(𝟏s=0+𝟏s=d2))).\left|\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}(v(x)-v(y))\cdot\nabla{\mathsf{g}}(x-y)d\left\lparen\frac{1}{N}\sum_{i=1}^{N}\delta_{x_{i}}-\mu\right\rparen^{\otimes 2}(x,y)\right|\\ \leq C\Bigg{(}\|\nabla v\|_{L^{\infty}}+\||\nabla|^{\frac{d-s}{2}}v\|_{L^{\frac{2d}{d-2-s}}}\mathbf{1}_{s<d-2}\Bigg{)}\Bigg{(}F_{N}(\underline{x}_{N},\mu)+CN^{-\frac{s+3}{(s+2)(s+1)}}\||\nabla|^{s+1-d}\mu\|_{L^{\infty}}\\ +C(1+\|\mu\|_{L^{\infty}})N^{-\frac{2}{(s+2)(s+1)}}\Big{(}1+(\log N)(\mathbf{1}_{s=0}+\mathbf{1}_{s=d-2})\Big{)}\Bigg{)}.

5.2. New estimates for globally superharmonic potentials

We now assume that r0=r_{0}=\infty in the condition iii, i.e. 𝗀{\mathsf{g}} is globally superharmonic. Under this more restrictive assumption, which holds in the model potential case (1.3), we can obtain versions of Proposition 5.1 and Proposition 5.4 that have a better balance of factors of μL\|\mu\|_{L^{\infty}} between terms. In particular, there is no η2\eta^{2} error term like there is in the right-hand side of inequality (5.8). This is important because if μt\mu^{t} is time-dependent and satisfies the decay bound (3.22), this term will contribute linear growth in time when integrated. As we shall see in Section 7.2, this better balance will be crucial to show that the error terms (i.e. those which are not FN(x¯Nt,μt)F_{N}(\underline{x}_{N}^{t},\mu^{t})) that result when estimating the right-hand side of (1.7) are integrable in time over the interval [0,)[0,\infty).

Proposition 5.8.

Let d3d\geq 3 and 0sd20\leq s\leq d-2. Assume that 𝗀{\mathsf{g}} is a potential satisfying conditions i, iii, iv, vi with r0=r_{0}=\infty. Suppose that x¯N(d)N\underline{x}_{N}\in({\mathbb{R}}^{d})^{N} is pairwise distinct and μ𝒫(d)L(d)\mu\in\mathcal{P}({\mathbb{R}}^{d})\cap L^{\infty}({\mathbb{R}}^{d}). In the case s=0s=0, also suppose that dlog(1+|x|)𝑑μ(x)<\int_{{\mathbb{R}}^{d}}\log(1+|x|)d\mu(x)<\infty. For any p>ds+2\infty\geq p>\frac{d}{s+2}, there exist constants C,Cp>0C,C_{p}>0 depending only on s,ds,d and the potential 𝗀{\mathsf{g}} through the assumed conditions, such that for every choice of 0<η1,,ηN<2dpd+2pd(p1)μL1d0<\eta_{1},\ldots,\eta_{N}<2^{-\frac{dp-d+2p}{d(p-1)}}\|\mu\|_{L^{\infty}}^{-\frac{1}{d}},

(5.20) 1N21ijN(𝗀(xjxi)𝗀ηi(xjxi))++C11Ni=1Nδxi(ηi)μH˙sd22FN(x¯N,μ)+CpμLγs,pNi=1Nηiλs,p(1+(|logηi|+|logμL|)𝟏s=0)+i=1NCηis(1+|logηi|𝟏s=0)N2,\frac{1}{N^{2}}\sum_{{1\leq i\neq j\leq N}}\left\lparen{\mathsf{g}}(x_{j}-x_{i})-{\mathsf{g}}_{\eta_{i}}(x_{j}-x_{i})\right\rparen_{+}+C^{-1}\left\|\frac{1}{N}\sum_{i=1}^{N}\delta_{x_{i}}^{(\eta_{i})}-\mu\right\|_{\dot{H}^{\frac{s-d}{2}}}^{2}\leq F_{N}(\underline{x}_{N},\mu)\\ +\frac{C_{p}\|\mu\|_{L^{\infty}}^{\gamma_{s,p}}}{N}\sum_{i=1}^{N}\eta_{i}^{\lambda_{s,p}}\left\lparen 1+\left\lparen|\log\eta_{i}|+|\log\|\mu\|_{L^{\infty}}|\right\rparen\mathbf{1}_{s=0}\right\rparen+\sum_{i=1}^{N}\frac{C\eta_{i}^{-s}(1+|\log\eta_{i}|\mathbf{1}_{s=0})}{N^{2}},

where the exponents γs,p,λs,p\gamma_{s,p},\lambda_{s,p} are defined by

(5.21) γs,p2p+spsdp+2pd,λs,p2p(ds)dp+2pd.\gamma_{s,p}\coloneqq\frac{2p+sp-s}{dp+2p-d},\qquad\lambda_{s,p}\coloneqq\frac{2p(d-s)}{dp+2p-d}.
Proof.

We modify the proof of [NRS21, Proposition 2.1]. Adding and subtracting δxi(ηi)\delta_{x_{i}}^{(\eta_{i})} and regrouping terms yields the decomposition

(5.22) FN(x¯N,μ)=(d)2𝗀(xy)d(1Ni=1Nδxi(ηi)μ)2(x,y)1N2i=1Nd𝗀ηi𝑑δ0(ηi)2Ni=1Nd(𝗀(yxi)𝗀ηi(yxi))𝑑μ(y)+1N21ijNd(𝗀(yxi)𝗀ηi(yxi))d(δxj+δxj(η))(y).\begin{split}F_{N}(\underline{x}_{N},\mu)&=\int_{({\mathbb{R}}^{d})^{2}}{\mathsf{g}}(x-y)d\left\lparen\frac{1}{N}\sum_{i=1}^{N}\delta_{x_{i}}^{(\eta_{i})}-\mu\right\rparen^{\otimes 2}(x,y)-\frac{1}{N^{2}}\sum_{i=1}^{N}\int_{{\mathbb{R}}^{d}}{\mathsf{g}}_{\eta_{i}}d\delta_{0}^{(\eta_{i})}\\ &\phantom{=}-\frac{2}{N}\sum_{i=1}^{N}\int_{{\mathbb{R}}^{d}}\left\lparen{\mathsf{g}}(y-x_{i})-{\mathsf{g}}_{\eta_{i}}(y-x_{i})\right\rparen d\mu(y)\\ &\phantom{=}+\frac{1}{N^{2}}\sum_{1\leq i\neq j\leq N}\int_{{\mathbb{R}}^{d}}\left\lparen{\mathsf{g}}(y-x_{i})-{\mathsf{g}}_{\eta_{i}}(y-x_{i})\right\rparen d(\delta_{x_{j}}+\delta_{x_{j}}^{(\eta)})(y).\end{split}

Since Δ𝗀0\Delta{\mathsf{g}}\leq 0 on d{\mathbb{R}}^{d} by assumption iii and δxj(ηj)\delta_{x_{j}}^{(\eta_{j})} is a positive measure, we have from (5.4) that

(5.23) 1N21ijNd(𝗀(yxi)𝗀ηi(yxi))d(δxj+δxj(ηj))(y)1N21ijN(𝗀(xjxi)𝗀ηi(xjxi))+.\begin{split}&\frac{1}{N^{2}}\sum_{1\leq i\neq j\leq N}\int_{{\mathbb{R}}^{d}}\left\lparen{\mathsf{g}}(y-x_{i})-{\mathsf{g}}_{\eta_{i}}(y-x_{i})\right\rparen d(\delta_{x_{j}}+\delta_{x_{j}}^{(\eta_{j})})(y)\\ &\geq\frac{1}{N^{2}}\sum_{{1\leq i\neq j\leq N}}\left\lparen{\mathsf{g}}(x_{j}-x_{i})-{\mathsf{g}}_{\eta_{i}}(x_{j}-x_{i})\right\rparen_{+}.\end{split}

Let R2ηR\geq 2\eta be a parameter to be specified shortly. Using assumption iv and the estimate (5.5), we find that

|d(𝗀(yxi)𝗀η(yxi))𝑑μ(y)|\displaystyle\left|\int_{{\mathbb{R}}^{d}}\left\lparen{\mathsf{g}}(y-x_{i})-{\mathsf{g}}_{\eta}(y-x_{i})\right\rparen d\mu(y)\right|
μL|yxi|R(|𝗀(yxi)|+|𝗀η(yxi)|)𝑑y+|yxi|>R|𝗀(yxi)𝗀η(yxi)|𝑑μ(y)\displaystyle\leq\|\mu\|_{L^{\infty}}\int_{|y-x_{i}|\leq R}\left\lparen|{\mathsf{g}}(y-x_{i})|+|{\mathsf{g}}_{\eta}(y-x_{i})|\right\rparen dy+\int_{|y-x_{i}|>R}\left|{\mathsf{g}}(y-x_{i})-{\mathsf{g}}_{\eta}(y-x_{i})\right|d\mu(y)
(5.24) CμL(Rds(1+|logR|𝟏s=0))+Cη2|yxi|>R|yxi|s2𝑑μ(y)\displaystyle\leq C\|\mu\|_{L^{\infty}}\left\lparen R^{d-s}(1+|\log R|\mathbf{1}_{s=0})\right\rparen+C\eta^{2}\int_{|y-x_{i}|>R}|y-x_{i}|^{-s-2}d\mu(y)

By Hölder’s inequality,

|yxi|>R|yxi|s2𝑑μ(y)\displaystyle\int_{|y-x_{i}|>R}|y-x_{i}|^{-s-2}d\mu(y) C(p(s+2)d)1pRdp(s+2)pμLp\displaystyle\leq C(p(s+2)-d)^{-\frac{1}{p}}R^{\frac{d-p(s+2)}{p}}\|\mu\|_{L^{p^{\prime}}}
(5.25) C(p(s+2)d)1pRdp(s+2)pμL1p\displaystyle\leq C(p(s+2)-d)^{-\frac{1}{p}}R^{\frac{d-p(s+2)}{p}}\|\mu\|_{L^{\infty}}^{\frac{1}{p}}

for any Hölder conjugate p,pp,p^{\prime} with ds+2<p\frac{d}{s+2}<p\leq\infty. Implicitly, we have used that μ\mu is a probability density, and the constant CC is independent of pp. Setting

(5.26) μLRds=η2Rdp(s+2)pμL1p,\|\mu\|_{L^{\infty}}R^{d-s}=\eta^{2}R^{\frac{d-p(s+2)}{p}}\|\mu\|_{L^{\infty}}^{\frac{1}{p}},

we get

(5.27) R=(η2μL1pp)pdpd+2p.R=\left\lparen\eta^{2}\|\mu\|_{L^{\infty}}^{\frac{1-p}{p}}\right\rparen^{\frac{p}{dp-d+2p}}.

In order for R2ηR\geq 2\eta, we need

(5.28) (η2μL1pp)pdpd+2p2ηη2dpd+2pd(p1)μL1d.\left\lparen\eta^{2}\|\mu\|_{L^{\infty}}^{\frac{1-p}{p}}\right\rparen^{\frac{p}{dp-d+2p}}\geq 2\eta\Longleftrightarrow\eta\leq 2^{-\frac{dp-d+2p}{d(p-1)}}\|\mu\|_{L^{\infty}}^{-\frac{1}{d}}.

Substituting in the above choice of RR, we obtain that

(5.29) 1Ni=1N|d(𝗀(yxi)𝗀ηi(yxi))𝑑μ(y)|CNμL2ps+spdpd+2p(p(s+2)d)1dp+2pd×i=1Nηi2p(ds)dp+2pd(1+(pdp+2pd|logηi2μL1pp|)𝟏s=0).\frac{1}{N}\sum_{i=1}^{N}\left|\int_{{\mathbb{R}}^{d}}\left\lparen{\mathsf{g}}(y-x_{i})-{\mathsf{g}}_{\eta_{i}}(y-x_{i})\right\rparen d\mu(y)\right|\leq\frac{C}{N}\|\mu\|_{L^{\infty}}^{\frac{2p-s+sp}{dp-d+2p}}\left\lparen p(s+2)-d\right\rparen^{-\frac{1}{dp+2p-d}}\\ \times\sum_{i=1}^{N}\eta_{i}^{\frac{2p(d-s)}{dp+2p-d}}\left\lparen 1+\left\lparen\frac{p}{dp+2p-d}\left|\log\eta_{i}^{2}\|\mu\|_{L^{\infty}}^{\frac{1-p}{p}}\right|\right\rparen\mathbf{1}_{s=0}\right\rparen.

Next, we use the relation (5.6) to bound

(5.30) 1N2i=1N|d𝗀ηi𝑑δ0(ηi)|i=1NCηis(1+|logηi|𝟏s=0)N2.\frac{1}{N^{2}}\sum_{i=1}^{N}\left|\int_{{\mathbb{R}}^{d}}{\mathsf{g}}_{\eta_{i}}d\delta_{0}^{(\eta_{i})}\right|\leq\sum_{i=1}^{N}\frac{C\eta_{i}^{-s}(1+|\log\eta_{i}|\mathbf{1}_{s=0})}{N^{2}}.

Collecting (5.23), (5.29), (5.30) and using the assumption vi with Plancherel’s theorem for the remaining term in (5.22), we arrive at the inequality in the statement of the proposition. ∎

Remark 5.9.

Evidently, the constant CpC_{p} in Proposition 5.8 blows up as pds+2+p\rightarrow{\frac{d}{s+2}}^{+}.

Remark 5.10.

Dropping the s,ps,p subscripts in γs,p,λs,p\gamma_{s,p},\lambda_{s,p}, we balance the error terms by setting

(5.31) CpμLγηiλ=ηisNηi=Cp1λ+sμLγλ+sN1λ+s,C_{p}\|\mu\|_{L^{\infty}}^{\gamma}\eta_{i}^{\lambda}=\frac{\eta_{i}^{-s}}{N}\Longleftrightarrow\eta_{i}=C_{p}^{-\frac{1}{\lambda+s}}\|\mu\|_{L^{\infty}}^{-\frac{\gamma}{\lambda+s}}N^{-\frac{1}{\lambda+s}},

which, for possibly larger constant Cp>0C_{p}>0, implies the lower bound

(5.32) FN(x¯N,μ)CpμLsdNλλ+s(1+(|logμL|+logN)𝟏s=0).F_{N}(\underline{x}_{N},\mu)\geq-C_{p}\|\mu\|_{L^{\infty}}^{\frac{s}{d}}N^{-\frac{\lambda}{\lambda+s}}\left\lparen 1+\left\lparen|\log\|\mu\|_{L^{\infty}}|+\log N\right\rparen\mathbf{1}_{s=0}\right\rparen.
Remark 5.11.

Just as in Remark 5.3, if instead of vi, we only assume that 𝗀^0\hat{{\mathsf{g}}}\geq 0 on d{0}{\mathbb{R}}^{d}\setminus\{0\}, then we have the bound

(5.33) 1N21ijN(𝗀(xjxi)𝗀ηi(xjxi))+FN(x¯N,μ)+i=1NCηis(1+|logηi|𝟏s=0)N2+CpμLγNi=1Nηiλ(1+(|logηi|+|logμL|)𝟏s=0).\frac{1}{N^{2}}\sum_{1\leq i\neq j\leq N}\left\lparen{\mathsf{g}}(x_{j}-x_{i})-{\mathsf{g}}_{\eta_{i}}(x_{j}-x_{i})\right\rparen_{+}\leq F_{N}(\underline{x}_{N},\mu)+\sum_{i=1}^{N}\frac{C\eta_{i}^{-s}(1+|\log\eta_{i}|\mathbf{1}_{s=0})}{N^{2}}\\ +\frac{C_{p}\|\mu\|_{L^{\infty}}^{\gamma}}{N}\sum_{i=1}^{N}\eta_{i}^{\lambda}\left\lparen 1+\left\lparen|\log\eta_{i}|+|\log\|\mu\|_{L^{\infty}}|\right\rparen\mathbf{1}_{s=0}\right\rparen.

Under the global superharmonicity assumption, we can also obtain a version of Proposition 5.4 without an η2\eta^{2} term and where every error term that is increasing in η\eta has a factor of μL\|\mu\|_{L^{\infty}}.

Proposition 5.12.

Let d3d\geq 3 and d2<s<dd-2<s<d. Suppose that x¯N(d)N\underline{x}_{N}\in({\mathbb{R}}^{d})^{N} is a pairwise distinct configuration and μ𝒫(d)L(d)\mu\in\mathcal{P}({\mathbb{R}}^{d})\cap L^{\infty}({\mathbb{R}}^{d}). Let 𝗀{\mathsf{g}} be a potential satisfying i, iii, iv with r0=r_{0}=\infty. Let 𝖦:d+m{0}{\mathsf{G}}:{\mathbb{R}}^{d+m}\setminus\{0\}\rightarrow{\mathbb{R}} be an extension 𝖦(x,0)=𝗀(x){\mathsf{G}}(x,0)={\mathsf{g}}(x) such that 𝖦{\mathsf{G}} satisfies conditions (1.14), (1.15), (1.16), (1.17) with r0=r_{0}=\infty. There exists a constant C>0C>0 only depending on s,ds,d and on 𝗀,𝖦{\mathsf{g}},{\mathsf{G}}, such that for every η1,,ηN>0\eta_{1},\ldots,\eta_{N}>0, we have

(5.34) 1N21ijN(𝗀(xjxi)𝖦ηi(xjxi,0))+FN(x¯N,μ)+CNi=1N(ηisN+μLηids).\frac{1}{N^{2}}\sum_{1\leq i\neq j\leq N}\left\lparen{\mathsf{g}}(x_{j}-x_{i})-{\mathsf{G}}_{\eta_{i}}(x_{j}-x_{i},0)\right\rparen_{+}\leq F_{N}(\underline{x}_{N},\mu)+\frac{C}{N}\sum_{i=1}^{N}\left\lparen\frac{\eta_{i}^{-s}}{N}+\|\mu\|_{L^{\infty}}\eta_{i}^{d-s}\right\rparen.
Proof.

We modify the proof of [NRS21, Proposition 2.2] in the same spirit as we did for Proposition 5.8. Adding and subtracting δXi(ηi)\delta_{X_{i}}^{(\eta_{i})} and regrouping terms, we find that

(5.35) FN(x¯N,μ)=(d+m)2𝖦(XY)d(1Ni=1NδXi(ηi)μ~)2(x,y)1N2i=1N(d+m)2𝖦(XY)d(δ0(ηi))2(X,Y)2Ni=1Nd+m(𝖦(YXi)𝖦ηi(YXi))𝑑μ~(Y)+1N21ijNd(𝖦(YXi)𝖦ηi(YXi))d(δXj+δXj(ηj))(Y).\begin{split}F_{N}(\underline{x}_{N},\mu)&=\int_{({\mathbb{R}}^{d+m})^{2}}{\mathsf{G}}(X-Y)d\left\lparen\frac{1}{N}\sum_{i=1}^{N}\delta_{X_{i}}^{(\eta_{i})}-\tilde{\mu}\right\rparen^{\otimes 2}(x,y)\\ &\phantom{=}-\frac{1}{N^{2}}\sum_{i=1}^{N}\int_{({\mathbb{R}}^{d+m})^{2}}{\mathsf{G}}(X-Y)d(\delta_{0}^{(\eta_{i})})^{\otimes 2}(X,Y)\\ &\phantom{=}-\frac{2}{N}\sum_{i=1}^{N}\int_{{\mathbb{R}}^{d+m}}\left\lparen{\mathsf{G}}(Y-X_{i})-{\mathsf{G}}_{\eta_{i}}(Y-X_{i})\right\rparen d\tilde{\mu}(Y)\\ &\phantom{=}+\frac{1}{N^{2}}\sum_{1\leq i\neq j\leq N}\int_{{\mathbb{R}}^{d}}\left\lparen{\mathsf{G}}(Y-X_{i})-{\mathsf{G}}_{\eta_{i}}(Y-X_{i})\right\rparen d({\delta}_{X_{j}}+{\delta}_{X_{j}}^{(\eta_{j})})(Y).\end{split}

By the inequality (5.13) and since δXi(ηi)\delta_{X_{i}}^{(\eta_{i})} is a positive measure in d+m{\mathbb{R}}^{d+m}, we have the lower bound

(5.36) 1N21ijNd(𝖦(YXi)𝖦ηi(YXi))d(δXj+δXj(ηj))(Y)1N21ijN(𝗀(xjxi)𝖦η(xjxi,0))+.\begin{split}&\frac{1}{N^{2}}\sum_{1\leq i\neq j\leq N}\int_{{\mathbb{R}}^{d}}\left\lparen{\mathsf{G}}(Y-X_{i})-{\mathsf{G}}_{\eta_{i}}(Y-X_{i})\right\rparen d({\delta}_{X_{j}}+{\delta}_{X_{j}}^{(\eta_{j})})(Y)\\ &\geq\frac{1}{N^{2}}\sum_{1\leq i\neq j\leq N}\left\lparen{\mathsf{g}}(x_{j}-x_{i})-{\mathsf{G}}_{\eta}(x_{j}-x_{i},0)\right\rparen_{+}.\end{split}

Letting R2ηR\geq 2\eta be a parameter to be determined, we have that

(5.37) d+m(𝖦(YXi)𝖦ηi(YXi))𝑑μ~(Y)=d(𝗀(yxi)𝖦ηi(yxi,0))𝑑μ(y)\displaystyle\int_{{\mathbb{R}}^{d+m}}\left\lparen{\mathsf{G}}(Y-X_{i})-{\mathsf{G}}_{\eta_{i}}(Y-X_{i})\right\rparen d\tilde{\mu}(Y)=\int_{{\mathbb{R}}^{d}}\left\lparen{\mathsf{g}}(y-x_{i})-{\mathsf{G}}_{\eta_{i}}(y-x_{i},0)\right\rparen d\mu(y)

by definition of μ~\tilde{\mu}. Since s>d2s>d-2 by assumption, we can use (5.14) and (1.16) to obtain the bound

(5.38) |yxi|R|𝗀(yxi)𝖦ηi((yxi,0))|𝑑μ(y)Cηi2Rds2μL.\int_{|y-x_{i}|\geq R}|{\mathsf{g}}(y-x_{i})-{\mathsf{G}}_{\eta_{i}}((y-x_{i},0))|d\mu(y)\leq C\eta_{i}^{2}R^{d-s-2}\|\mu\|_{L^{\infty}}.

Using 1.16 to estimate directly the integral over |yxi|<R|y-x_{i}|<R as in (5.24), it follows that

(5.39) |d+m(𝗀(yxi)𝖦ηi(yxi,0))𝑑μ(y)|CμL(Rds+ηi2Rds2).\left|\int_{{\mathbb{R}}^{d+m}}\left\lparen{\mathsf{g}}(y-x_{i})-{\mathsf{G}}_{\eta_{i}}(y-x_{i},0)\right\rparen d\mu(y)\right|\leq C\|\mu\|_{L^{\infty}}\left\lparen R^{d-s}+\eta_{i}^{2}R^{d-s-2}\right\rparen.

We can them optimize the choice of RR by setting R=2ηiR=2\eta_{i}. After a little bookkeeping, we have shown that

(5.40) 1Ni=1N|d+m(𝗀(yxi)𝖦ηi(yxi,0))𝑑μ(y)|CηidsμL.\frac{1}{N}\sum_{i=1}^{N}\left|\int_{{\mathbb{R}}^{d+m}}\left\lparen{\mathsf{g}}(y-x_{i})-{\mathsf{G}}_{\eta_{i}}(y-x_{i},0)\right\rparen d\mu(y)\right|\leq C\eta_{i}^{d-s}\|\mu\|_{L^{\infty}}.

Finally, using the relation (5.15), we have the self-interaction bound

(5.41) 1N2i=1N|(d+m)2𝖦(XY)d(δ0(ηi))2(X,Y)|CηsN\frac{1}{N^{2}}\sum_{i=1}^{N}\left|\int_{({\mathbb{R}}^{d+m})^{2}}{\mathsf{G}}(X-Y)d(\delta_{0}^{(\eta_{i})})^{\otimes 2}(X,Y)\right|\leq\frac{C\eta^{-s}}{N}

and using assumption 1.17 with Plancherel’s theorem, we have the lower bound

(5.42) (d+m)2𝖦(XY)d(1Ni=1Nδxi(ηi)μ~)(X,Y)0.\int_{({\mathbb{R}}^{d+m})^{2}}{\mathsf{G}}(X-Y)d\left\lparen\frac{1}{N}\sum_{i=1}^{N}\delta_{x_{i}}^{(\eta_{i})}-\tilde{\mu}\right\rparen(X,Y)\geq 0.

Combining these observations with (5.36), (5.40), we arrive at the inequality in the statement of the proposition. ∎

Remark 5.13.

Similar to Remark 5.10, we can balance the error terms in (5.34) by choosing ηi=(μLN)1d\eta_{i}=(\|\mu\|_{L^{\infty}}N)^{-\frac{1}{d}}, which implies the lower bound

(5.43) FN(x¯N,μ)CμLsdNdsd.F_{N}(\underline{x}_{N},\mu)\geq-C\|\mu\|_{L^{\infty}}^{\frac{s}{d}}N^{-\frac{d-s}{d}}.

Analogous to 5.6, we can use assumption viii for admissible potentials 𝗀{\mathsf{g}} together with Proposition 5.8 (if 0sd40\leq s\leq d-4) and Proposition 5.12 (if d4<s<d2d-4<s<d-2) to obtain the following result.

Corollary 5.14.

Let d3d\geq 3 and 0s<d20\leq s<d-2. Suppose that x¯N(d)N\underline{x}_{N}\in({\mathbb{R}}^{d})^{N} is pairwise distinct and μ𝒫(d)L(d)\mu\in\mathcal{P}({\mathbb{R}}^{d})\cap L^{\infty}({\mathbb{R}}^{d}). Let 𝗀{\mathsf{g}} be a potential satisfying assumptions i, iii, iv, vi, viii with r0=r_{0}=\infty. There exists a constant C>0C>0 depending only on s,ds,d and 𝗀,𝖦{\mathsf{g}},{\mathsf{G}}, such that the following holds. If 0sd40\leq s\leq d-4, then for any p>ds+4\infty\geq p>\frac{d}{s+4}, CC also depends on pp and

(5.44) (d)2(Δ𝗀)(xy)d(1Ni=1Nδxiμ)2(x,y)CpμLs+2dNλs+2,pλs+2,p+s+2.\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}(-\Delta{\mathsf{g}})(x-y)d\left\lparen\frac{1}{N}\sum_{i=1}^{N}\delta_{x_{i}}-\mu\right\rparen^{\otimes 2}(x,y)\geq-C_{p}\|\mu\|_{L^{\infty}}^{\frac{s+2}{d}}N^{-\frac{\lambda_{s+2,p}}{\lambda_{s+2,p}+s+2}}.

where λs+2,p\lambda_{s+2,p} is as defined in (5.21). If d4<s<d2d-4<s<d-2, then

(5.45) (d)2(Δ𝗀)(xy)d(1Ni=1Nδxiμ)2(x,y)CμLs+2dNds2d.\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}(-\Delta{\mathsf{g}})(x-y)d\left\lparen\frac{1}{N}\sum_{i=1}^{N}\delta_{x_{i}}-\mu\right\rparen^{\otimes 2}(x,y)\geq-C\|\mu\|_{L^{\infty}}^{\frac{s+2}{d}}N^{-\frac{d-s-2}{d}}.
Proof.

If 0sd40\leq s\leq d-4, then we use (5.33) with ss replaced by s+2s+2 and choosing
ηi=μLγs+2,pλs+2,p+s+2N1λs+2,p+s+2\eta_{i}=\|\mu\|_{L^{\infty}}^{-\frac{\gamma_{s+2,p}}{\lambda_{s+2,p}+s+2}}N^{-\frac{1}{\lambda_{s+2,p}+s+2}}. If d4<s<d2d-4<s<d-2, then we use (5.34) with ss replaced by s+2s+2 and choosing ηi=(μLN)1d\eta_{i}=(\|\mu\|_{L^{\infty}}N)^{-\frac{1}{d}}. ∎

Repeating the proof of [NRS21, Proposition 4.1], except now using Proposition 5.8 instead of Proposition 5.1, we can obtain a renormalized commutator estimate (cf. Proposition 5.7) with better distribution of norms of μ\mu.

Proposition 5.15.

Let d3d\geq 3 and 0sd20\leq s\leq d-2. Let x¯N(d)N\underline{x}_{N}\in({\mathbb{R}}^{d})^{N} be pairwise distinct, and μ𝒫(d)L(d)\mu\in\mathcal{P}({\mathbb{R}}^{d})\cap L^{\infty}({\mathbb{R}}^{d}). If s=0s=0, assume that dlog(1+|x|)𝑑μ(x)<\int_{{\mathbb{R}}^{d}}\log(1+|x|)d\mu(x)<\infty. Let vv be a continuous vector field on d{\mathbb{R}}^{d}. For every p>ds+2\infty\geq p>\frac{d}{s+2}, there exists a constant CpC_{p} depending only d,sd,s and on the potential 𝗀{\mathsf{g}} through assumptions ivii, ix such that for all N>(2dpd+p(p1)μL)(s+1)(s+λs,p)dN>(2^{\frac{dp-d+p}{(p-1)}}\|\mu\|_{L^{\infty}})^{\frac{(s+1)(s+\lambda_{s,p})}{d}},

(5.46) |(d)2(v(x)v(y))𝗀(xy)d(1Ni=1Nδxiμ)2(x,y)|CvL||s+1dμLNs+1+λs,p(s+λs,p)(1+s)+C(vL+||ds2vL2dd2s𝟏s<d2)(FN(x¯N,μ)+Cp(1+μLγs,p)Nλs,p(s+λs,p)(1+s)(1+(logN+|logμL|)𝟏s=0)),\left|\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}(v(x)-v(y))\cdot\nabla{\mathsf{g}}(x-y)d\left\lparen\frac{1}{N}\sum_{i=1}^{N}\delta_{x_{i}}-\mu\right\rparen^{\otimes 2}(x,y)\right|\\ \leq C\|\nabla v\|_{L^{\infty}}\||\nabla|^{s+1-d}\mu\|_{L^{\infty}}N^{-\frac{s+1+\lambda_{s,p}}{(s+\lambda_{s,p})(1+s)}}+C\Bigg{(}\|\nabla v\|_{L^{\infty}}+\||\nabla|^{\frac{d-s}{2}}v\|_{L^{\frac{2d}{d-2-s}}}\mathbf{1}_{s<d-2}\Bigg{)}\Bigg{(}F_{N}(\underline{x}_{N},\mu)\\ +C_{p}\left\lparen 1+\|\mu\|_{L^{\infty}}^{\gamma_{s,p}}\right\rparen N^{-\frac{\lambda_{s,p}}{(s+\lambda_{s,p})(1+s)}}\left\lparen 1+\left\lparen\log N+|\log\|\mu\|_{L^{\infty}}|\right\rparen\mathbf{1}_{s=0}\right\rparen\Bigg{)},

where γs,p,λs,p\gamma_{s,p},\lambda_{s,p} are as in (5.21).

Proof.

Since s,ps,p are fixed, we drop the subscripts in γs,p,λs,p\gamma_{s,p},\lambda_{s,p}. Repeating the steps in the proof of [NRS21, Proposition 4.1], we find that the left-hand side of (5.46) is controlled by

(5.47) C(vL+||ds2vL2dd2s𝟏s<d2)(FN(x¯N,μ)+CpμLγηλ(1+(|logη|+|logμL|)𝟏s=0)+Cηs(1+|logη|𝟏s=0)N)+CvL(εs(1+|logε|𝟏s=0)N+CpμLγελ(1+(|logε|+|logμL|)𝟏s=0)+η||s+1dμL+ηεs+1),C\Bigg{(}\|\nabla v\|_{L^{\infty}}+\||\nabla|^{\frac{d-s}{2}}v\|_{L^{\frac{2d}{d-2-s}}}\mathbf{1}_{s<d-2}\Bigg{)}\Bigg{(}F_{N}(\underline{x}_{N},\mu)\\ +C_{p}\|\mu\|_{L^{\infty}}^{\gamma}\eta^{\lambda}\left\lparen 1+\left\lparen|\log\eta|+|\log\|\mu\|_{L^{\infty}}|\right\rparen\mathbf{1}_{s=0}\right\rparen+\frac{C\eta^{-s}(1+|\log\eta|\mathbf{1}_{s=0})}{N}\Bigg{)}\\ +C\|\nabla v\|_{L^{\infty}}\Bigg{(}\frac{\varepsilon^{-s}(1+|\log\varepsilon|\mathbf{1}_{s=0})}{N}+C_{p}\|\mu\|_{L^{\infty}}^{\gamma}\varepsilon^{\lambda}\left\lparen 1+\left\lparen|\log\varepsilon|+|\log\|\mu\|_{L^{\infty}}|\right\rparen\mathbf{1}_{s=0}\right\rparen\\ +\eta\||\nabla|^{s+1-d}\mu\|_{L^{\infty}}+\frac{\eta}{\varepsilon^{s+1}}\Bigg{)},

where 0<2ηε<2dpd+2pd(p1)μL1d0<2\eta\leq\varepsilon<2^{-\frac{dp-d+2p}{d(p-1)}}\|\mu\|_{L^{\infty}}^{-\frac{1}{d}} and p>ds+2\infty\geq p>\frac{d}{s+2}. Since 0sd20\leq s\leq d-2 by assumption, we balance error terms (i.e. those terms which are not FN(x¯N,μ)F_{N}(\underline{x}_{N},\mu)) by setting

(5.48) ηεs+1=ηsN=ελ,\frac{\eta}{\varepsilon^{s+1}}=\frac{\eta^{-s}}{N}=\varepsilon^{\lambda},

which yields η=ελ+s+1\eta=\varepsilon^{\lambda+s+1} and ε=N1(1+s)(s+λ)\varepsilon=N^{-\frac{1}{(1+s)(s+\lambda)}}. To ensure that ε<2dpd+2pd(p1)μL1d\varepsilon<2^{-\frac{dp-d+2p}{d(p-1)}}\|\mu\|_{L^{\infty}}^{-\frac{1}{d}}, we require that

(5.49) N1(1+s)(s+λ)<2dpd+2pd(p1)μL1d2(dpd+2p)(1+s)(s+λ)d(p1)μL(1+s)(s+λ)d<N.N^{-\frac{1}{(1+s)(s+\lambda)}}<2^{-\frac{dp-d+2p}{d(p-1)}}\|\mu\|_{L^{\infty}}^{-\frac{1}{d}}\Longleftrightarrow 2^{\frac{(dp-d+2p)(1+s)(s+\lambda)}{d(p-1)}}\|\mu\|_{L^{\infty}}^{\frac{(1+s)(s+\lambda)}{d}}<N.

Substituting these choices back into (5.47), we arrive at the inequality in the statement of the proposition. ∎

6. Evolution of the modulated energy

Our next task is to rigorously compute the time-derivative of the modulated energy, which we recall is a real-valued stochastic process. Since the potential 𝗀{\mathsf{g}} is not C2C^{2} due to its singularity at the origin, we cannot directly apply Itô’s lemma to FN(x¯Nt,μt)F_{N}(\underline{x}_{N}^{t},\mu^{t}), as we formally did in the introduction to obtain (1.7). Instead, we proceed by a truncation and stopping time argument, similar to that used Section 4 to prove the well-posedness of the NN-body dynamics.

We define the truncated modulated energy

(6.1) FN,ε(x¯N,εt,μt)(d)2𝗀(ε)(xy)d(μN,εtμt)2(x,y),F_{N,\varepsilon}(\underline{x}_{N,\varepsilon}^{t},\mu^{t})\coloneqq\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}{\mathsf{g}}_{(\varepsilon)}(x-y)d(\mu_{N,\varepsilon}^{t}-\mu^{t})^{\otimes 2}(x,y),

where 𝗀(ε){\mathsf{g}}_{(\varepsilon)} is as defined in (4.2), x¯N,ε\underline{x}_{N,\varepsilon} is the solution to the truncated system (4.5), and μN,ε\mu_{N,\varepsilon} denotes the empirical measure induced by x¯N,ε\underline{x}_{N,\varepsilon}. Comparing this expression to the definition of FN(x¯N,μ)F_{N}(\underline{x}_{N},\mu) above, we have just replaced the potential 𝗀{\mathsf{g}} with the potential 𝗀(ε){\mathsf{g}}_{(\varepsilon)} and replaced x¯N\underline{x}_{N} with x¯N,ε\underline{x}_{N,\varepsilon}. Thanks to the regularity of the truncated potential 𝗀(ε){\mathsf{g}}_{(\varepsilon)}, we can rigorously apply Itô’s lemma to FN,ε(x¯N,εt,μt)F_{N,\varepsilon}(\underline{x}_{N,\varepsilon}^{t},\mu^{t}).

Lemma 6.1.

Let x¯N,ε\underline{x}_{N,\varepsilon} be a solution to the system (4.5), and let μC([0,);L1(d)L(d))\mu\in C([0,\infty);L^{1}({\mathbb{R}}^{d})\cap L^{\infty}({\mathbb{R}}^{d})) be a solution to equation (1.5). Then for every choice of ε\varepsilon satisfying

(6.2) 0<ε12min1ijN|xi0xj0|,0<\varepsilon\leq\frac{1}{2}\min_{1\leq i\neq j\leq N}|x_{i}^{0}-x_{j}^{0}|,

it holds with probability one that for all t0t\geq 0,

(6.3) FN,ε(x¯N,εt,μt)=FN,ε(x¯N0,μ0)+2N31i,j,kNk,ji0t𝗀(ε)(xi,εκxj,εκ)𝕄𝗀(ε)(xi,εκxk,εκ)𝑑κ+2Ni=1N0t(𝗀(ε)div(uκμκ))(xi,εκ)𝑑κ+2N21ijN0t(𝗀(ε)μκ)(xi,εκ)𝕄𝗀(ε)(xi,εκxj,εκ)𝑑κ20t𝗀(ε)div(uκμκ),μκL2𝑑κ+2σ0t(2)2Δ𝗀(ε)(xy)d(μN,εκμκ)2(x,y)𝑑κ+22σNi=1N0td{xi,εκ}𝗀(ε)(xi,εκy)d(μN,εκμκ)(y)𝑑Wiκ,F_{N,\varepsilon}(\underline{x}_{N,\varepsilon}^{t},\mu^{t})=F_{N,\varepsilon}(\underline{x}_{N}^{0},\mu^{0})+\frac{2}{N^{3}}\sum_{\begin{subarray}{c}1\leq i,j,k\leq N\\ k,j\neq i\end{subarray}}\int_{0}^{t}\nabla{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}^{\kappa}-x_{j,\varepsilon}^{\kappa})\cdot{\mathbb{M}}\nabla{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}^{\kappa}-x_{k,\varepsilon}^{\kappa})d\kappa\\ +\frac{2}{N}\sum_{i=1}^{N}\int_{0}^{t}({\mathsf{g}}_{(\varepsilon)}\ast\operatorname{\mathrm{div}}(u^{\kappa}\mu^{\kappa}))(x_{i,\varepsilon}^{\kappa})d\kappa+\frac{2}{N^{2}}\sum_{1\leq i\neq j\leq N}\int_{0}^{t}(\nabla{\mathsf{g}}_{(\varepsilon)}\ast\mu^{\kappa})(x_{i,\varepsilon}^{\kappa})\cdot{\mathbb{M}}\nabla{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}^{\kappa}-x_{j,\varepsilon}^{\kappa})d\kappa\\ -2\int_{0}^{t}\langle{\mathsf{g}}_{(\varepsilon)}\ast\operatorname{\mathrm{div}}(u^{\kappa}\mu^{\kappa}),\mu^{\kappa}\rangle_{L^{2}}d\kappa+2\sigma\int_{0}^{t}\int_{({\mathbb{R}}^{2})^{2}\setminus\triangle}\Delta{\mathsf{g}}_{(\varepsilon)}(x-y)d(\mu_{N,\varepsilon}^{\kappa}-\mu^{\kappa})^{\otimes 2}(x,y)d\kappa\\ +\frac{2\sqrt{2\sigma}}{N}\sum_{i=1}^{N}\int_{0}^{t}\int_{{\mathbb{R}}^{d}\setminus\{x_{i,\varepsilon}^{\kappa}\}}\nabla{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}^{\kappa}-y)d(\mu_{N,\varepsilon}^{\kappa}-\mu^{\kappa})(y)\cdot dW_{i}^{\kappa},

where μN,ε1Ni=1Nδxi,ϵ\mu_{N,\varepsilon}\coloneqq\frac{1}{N}\sum_{i=1}^{N}\delta_{x_{i,\epsilon}} and u𝕄𝗀μu\coloneqq{\mathbb{M}}\nabla{\mathsf{g}}\ast\mu.

Proof.

By approximation, we may assume without loss of generality that μ\mu is smooth and rapidly decaying at infinity. We split the modulated energy into a sum of three terms, defined by

(6.4) Term1\displaystyle\mathrm{Term}_{1} 1N21ijN𝗀(ε)(xi,εxj,ε),\displaystyle\coloneqq\frac{1}{N^{2}}\sum_{1\leq i\neq j\leq N}{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}-x_{j,\varepsilon}),
(6.5) Term2\displaystyle\mathrm{Term}_{2} 2Ni=1N2𝗀(ε)(xi,εy)𝑑μ(y)=2Ni=1N𝗀(ε)μ(xi,ε),\displaystyle\coloneqq-\frac{2}{N}\sum_{i=1}^{N}\int_{{\mathbb{R}}^{2}}{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}-y)d\mu(y)=-\frac{2}{N}\sum_{i=1}^{N}{\mathsf{g}}_{(\varepsilon)}\ast\mu(x_{i,\varepsilon}),
(6.6) Term3\displaystyle\mathrm{Term}_{3} (2)2𝗀(ε)(xy)𝑑μ2(x,y)=𝗀(ε)μ,μL2,\displaystyle\coloneqq\int_{({\mathbb{R}}^{2})^{2}}{\mathsf{g}}_{(\varepsilon)}(x-y)d\mu^{\otimes 2}(x,y)=\langle{\mathsf{g}}_{(\varepsilon)}\ast\mu,\mu\rangle_{L^{2}},

and compute the stochastic/deterministic differential equation satisfied by each of these terms. Of course, Term1,,Term3\mathrm{Term}_{1},\ldots,\mathrm{Term}_{3} depend on ε\varepsilon, but since ε\varepsilon is fixed, we omit this dependence.

Term1\mathrm{Term}_{1}:

By Itô’s lemma and Remark 4.1, we have, for iji\neq j, that 𝗀(ε)(xi,εxj,ε){\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}-x_{j,\varepsilon}) satisfies the SDE

d𝗀(ε)(xi,εxj,ε)\displaystyle d{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}-x_{j,\varepsilon}) =𝗀(ε)(xi,εxj,ε)d(xi,εxj,ε)+122𝗀(ε)(xi,εxj,ε):d[xi,εxj,ε]\displaystyle=\nabla{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}-x_{j,\varepsilon})\cdot d(x_{i,\varepsilon}-x_{j,\varepsilon})+\frac{1}{2}\nabla^{\otimes 2}{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}-x_{j,\varepsilon}):d[x_{i,\varepsilon}-x_{j,\varepsilon}]
=𝗀(ε)(xi,εxj,ε)(1N1kNki𝕄𝗀(ε)(xi,εxk,ε)1N1kNkj𝕄𝗀(ε)(xj,εxk,ε))dt\displaystyle=\nabla{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}-x_{j,\varepsilon})\cdot\left\lparen\frac{1}{N}\sum_{\begin{subarray}{c}1\leq k\leq N\\ k\neq i\end{subarray}}{\mathbb{M}}\nabla{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}-x_{k,\varepsilon})-\frac{1}{N}\sum_{\begin{subarray}{c}1\leq k\leq N\\ k\neq j\end{subarray}}{\mathbb{M}}\nabla{\mathsf{g}}_{(\varepsilon)}(x_{j,\varepsilon}-x_{k,\varepsilon})\right\rparen dt
(6.7) +2σ𝗀(ε)(xi,εxj,ε)d(WiWj)+2σ(2𝗀(ε)(xi,εxj,ε):𝕀)dt.\displaystyle\phantom{=}+\sqrt{2\sigma}\nabla{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}-x_{j,\varepsilon})\cdot d(W_{i}-W_{j})+2\sigma(\nabla^{\otimes 2}{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}-x_{j,\varepsilon}):\mathbb{I})dt.

Evidently,

(6.8) 2σ2𝗀(ε)(xi,εxj,ε):𝕀=2σΔ𝗀(ε)(xi,εxj,ε)2\sigma\nabla^{\otimes 2}{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}-x_{j,\varepsilon}):\mathbb{I}=2\sigma\Delta{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}-x_{j,\varepsilon})

Thus by symmetry under swapping iji\leftrightarrow j, and after integrating in time, we obtain

(6.9) Term1(t)=1N21ijN𝗀(xi0xj0)+22σN21ijN0t𝗀(ε)(xi,εκxj,εκ)𝑑Wiκ+2N31ijN1kNki0t𝗀(ε)(xi,εκxj,εκ)𝕄𝗀(ε)(xi,εκxk,εκ)𝑑κ+2σ1ijN0tΔ𝗀(ε)(xi,εκxj,εκ)𝑑κ,\mathrm{Term}_{1}(t)=\frac{1}{N^{2}}\sum_{1\leq i\neq j\leq N}{\mathsf{g}}(x_{i}^{0}-x_{j}^{0})+\frac{2\sqrt{2\sigma}}{N^{2}}\sum_{1\leq i\neq j\leq N}\int_{0}^{t}\nabla{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}^{\kappa}-x_{j,\varepsilon}^{\kappa})\cdot dW_{i}^{\kappa}\\ +\frac{2}{N^{3}}\sum_{1\leq i\neq j\leq N}\sum_{\begin{subarray}{c}1\leq k\leq N\\ k\neq i\end{subarray}}\int_{0}^{t}\nabla{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}^{\kappa}-x_{j,\varepsilon}^{\kappa})\cdot{\mathbb{M}}\nabla{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}^{\kappa}-x_{k,\varepsilon}^{\kappa})d\kappa\\ +2\sigma\sum_{1\leq i\neq j\leq N}\int_{0}^{t}\Delta{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}^{\kappa}-x_{j,\varepsilon}^{\kappa})d\kappa,

provided that ε12minij|xi,0xj,0|\varepsilon\leq\frac{1}{2}\min_{i\neq j}|x_{i,0}-x_{j,0}|.

Term2\mathrm{Term}_{2}:

Defining f(t,x)(𝗀(ε)μt)(x)f(t,x)\coloneqq({\mathsf{g}}_{(\varepsilon)}\ast\mu^{t})(x), we first observe from equation (1.5) that

(6.10) tf=𝗀(ε)(div(μu)+σΔμ)),{\partial}_{t}f={\mathsf{g}}_{(\varepsilon)}\ast\left\lparen-\operatorname{\mathrm{div}}(\mu u)+\sigma\Delta\mu)\right\rparen,

Applying Itô’s lemma with the time-dependent function ff, we find that

df(t,xi,ε)\displaystyle df(t,x_{i,\varepsilon}) =tf(t,xi,ε)dt+f(t,xi,ε)dxi,ε+122f(t,xi,ε):d[xi,ε]\displaystyle={\partial}_{t}f(t,x_{i,\varepsilon})dt+\nabla f(t,x_{i,\varepsilon})\cdot dx_{i,\varepsilon}+\frac{1}{2}\nabla^{\otimes 2}f(t,x_{i,\varepsilon}):d[x_{i,\varepsilon}]
=𝗀(ε)div(utμt)(xi,ε)dt+σ(𝗀(ε)Δμt)(xi,ε)dt\displaystyle=-{\mathsf{g}}_{(\varepsilon)}\ast\operatorname{\mathrm{div}}(u^{t}\mu^{t})(x_{i,\varepsilon})dt+\sigma({\mathsf{g}}_{(\varepsilon)}\ast\Delta\mu^{t})(x_{i,\varepsilon})dt
+(𝗀(ε)μt)(xi,ε)1N1kNki𝕄𝗀(ε)(xi,εxk,ε)dt\displaystyle\phantom{=}+\nabla({\mathsf{g}}_{(\varepsilon)}\ast\mu^{t})(x_{i,\varepsilon})\cdot\frac{1}{N}\sum_{\begin{subarray}{c}1\leq k\leq N\\ k\neq i\end{subarray}}{\mathbb{M}}\nabla{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}-x_{k,\varepsilon})dt
(6.11) +2σ(𝗀(ε)μt)(xi,ε)dWi+σ(2(𝗀(ε)μt)(xi,ε):𝕀)dt,\displaystyle\phantom{=}+\sqrt{2\sigma}\nabla({\mathsf{g}}_{(\varepsilon)}\ast\mu^{t})(x_{i,\varepsilon})\cdot dW_{i}+\sigma(\nabla^{\otimes 2}({\mathsf{g}}_{(\varepsilon)}\ast\mu^{t})(x_{i,\varepsilon}):\mathbb{I})dt,

where we also use Remark 4.1 to obtain the ultimate line. Noting that

(6.12) σ2(𝗀(ε)μt)(xi,ε):𝕀=σΔ(𝗀(ε)μt)(xi,ε),\sigma\nabla^{\otimes 2}({\mathsf{g}}_{(\varepsilon)}\ast\mu^{t})(x_{i,\varepsilon}):\mathbb{I}=\sigma\Delta({\mathsf{g}}_{(\varepsilon)}\ast\mu^{t})(x_{i,\varepsilon}),

we conclude that

(6.13) Term2(t)=2N𝗀(ε)μ0(xi0)+2Ni=1N0t𝗀(ε)div(uκμκ)(xi,εκ)𝑑κ2σNi=1N0t(𝗀(ε)Δμκ)(xi,εκ)𝑑κ2N21ikN0t(𝗀(ε)μκ)(xi,εκ)𝕄𝗀(ε)(xi,εκxk,εκ)𝑑κ22σNi=1N0t(𝗀(ε)μκ)(xi,εκ)𝑑Wiκ.\mathrm{Term}_{2}(t)=-\frac{2}{N}{\mathsf{g}}_{(\varepsilon)}\ast\mu^{0}(x_{i}^{0})+\frac{2}{N}\sum_{i=1}^{N}\int_{0}^{t}{\mathsf{g}}_{(\varepsilon)}\ast\operatorname{\mathrm{div}}(u^{\kappa}\mu^{\kappa})(x_{i,\varepsilon}^{\kappa})d\kappa-\frac{2\sigma}{N}\sum_{i=1}^{N}\int_{0}^{t}({\mathsf{g}}_{(\varepsilon)}\ast\Delta\mu^{\kappa})(x_{i,\varepsilon}^{\kappa})d\kappa\\ -\frac{2}{N^{2}}\sum_{1\leq i\neq k\leq N}\int_{0}^{t}\nabla({\mathsf{g}}_{(\varepsilon)}\ast\mu^{\kappa})(x_{i,\varepsilon}^{\kappa})\cdot{\mathbb{M}}\nabla{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}^{\kappa}-x_{k,\varepsilon}^{\kappa})d\kappa-\frac{2\sqrt{2\sigma}}{N}\sum_{i=1}^{N}\int_{0}^{t}\nabla({\mathsf{g}}_{(\varepsilon)}\ast\mu^{\kappa})(x_{i,\varepsilon}^{\kappa})\cdot dW_{i}^{\kappa}.
Term3\mathrm{Term}_{3}:

Using equation (6.10) and symmetry under swapping xyx\leftrightarrow y, we find that

(6.14) Term3(t)=𝗀εμ0,μ0+20t𝗀(ε)(div(uκμκ)+σΔμκ),μκL2𝑑κ.\begin{split}\mathrm{Term}_{3}(t)&=\langle{\mathsf{g}}_{\varepsilon}\ast\mu^{0},\mu^{0}\rangle+2\int_{0}^{t}\langle{\mathsf{g}}_{(\varepsilon)}\ast(-\operatorname{\mathrm{div}}(u^{\kappa}\mu^{\kappa})+\sigma\Delta\mu^{\kappa}),\mu^{\kappa}\rangle_{L^{2}}d\kappa.\end{split}

Combining the identities (6.9), (6.13) and (6.14) completes the proof of the lemma. ∎

We now proceed to group terms following the proof of [Ser20, Lemma 2.1]. We leave filling in the details to the reader, as it requires nothing new from the aforementioned work.

Lemma 6.2.

Let x¯N,ε\underline{x}_{N,\varepsilon} and μ\mu be as in Lemma 6.1. Then for every ε\varepsilon satisfying (6.2), it holds with probability 1 that for all t0t\geq 0,

(6.15) FN,ε(x¯N,εt,μt)FN,ε(x¯N0,μ0)0t(d)2(uεκ(x)uεκ(y))𝗀(ε)(xy)d(μN,εκμκ)2(x,y)+2σ0t(d)2Δ𝗀(ε)(xy)d(μN,εκμκ)2(x,y)𝑑κ+2Ni=1N0t(𝗀(ε)div((uκuεκ)μκ))(xi,εκ)𝑑κ20t𝗀(ε)div((uκuεκ)μκ),μκL2𝑑κ+22σNi=1N0tP.V.d{xi,εκ}𝗀(ε)(xi,εκy)d(μN,εκμκ)(y)𝑑Wiκ,F_{N,\varepsilon}(\underline{x}_{N,\varepsilon}^{t},\mu^{t})-F_{N,\varepsilon}(\underline{x}_{N}^{0},\mu^{0})\leq\int_{0}^{t}\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}\left\lparen u_{\varepsilon}^{\kappa}(x)-u_{\varepsilon}^{\kappa}(y)\right\rparen\cdot\nabla{\mathsf{g}}_{(\varepsilon)}(x-y)d(\mu_{N,\varepsilon}^{\kappa}-\mu^{\kappa})^{\otimes 2}(x,y)\\ +2\sigma\int_{0}^{t}\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}\Delta{\mathsf{g}}_{(\varepsilon)}(x-y)d(\mu_{N,\varepsilon}^{\kappa}-\mu^{\kappa})^{\otimes 2}(x,y)d\kappa+\frac{2}{N}\sum_{i=1}^{N}\int_{0}^{t}({\mathsf{g}}_{(\varepsilon)}\ast\operatorname{\mathrm{div}}((u^{\kappa}-u_{\varepsilon}^{\kappa})\mu^{\kappa}))(x_{i,\varepsilon}^{\kappa})d\kappa\\ -2\int_{0}^{t}\langle{\mathsf{g}}_{(\varepsilon)}\ast\operatorname{\mathrm{div}}((u^{\kappa}-u_{\varepsilon}^{\kappa})\mu^{\kappa}),\mu^{\kappa}\rangle_{L^{2}}d\kappa+\frac{2\sqrt{2\sigma}}{N}\sum_{i=1}^{N}\int_{0}^{t}\PV\int_{{\mathbb{R}}^{d}\setminus\{x_{i,\varepsilon}^{\kappa}\}}\nabla{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}^{\kappa}-y)d(\mu_{N,\varepsilon}^{\kappa}-\mu^{\kappa})(y)\cdot dW_{i}^{\kappa},

where uε𝕄𝗀(ε)μu_{\varepsilon}\coloneqq{\mathbb{M}}\nabla{\mathsf{g}}_{(\varepsilon)}\ast\mu.

We are now prepared to remove the truncation by passing to the limit ε0+\varepsilon\rightarrow 0^{+}. To this end, we recall from Section 4 the stopping time τε\tau_{\varepsilon} defined in (4.7). From Proposition 4.5, we know that limε0τε=\lim_{\varepsilon\to 0}\tau_{\varepsilon}=\infty a.s. The next proposition, the culmination of our work so far, is the main result of this subsection. It gives a functional inequality for the expected magnitude of the modulated energy, which serves as the first step in our Gronwall argument, and should be interpreted as the “rigorous version” of the inequality (1.7) from the introduction.

Proposition 6.3.

For all t0t\geq 0, we have the inequality

(6.16) 𝔼(FN(x¯Nt,μt)FN(x¯N0,μ0))2σ𝔼(0t(d)2Δ𝗀(xy)d(μNκμκ)2(x,y)𝑑κ)+𝔼(0t|(d)2(uκ(x)uκ(y))𝗀(xy)d(μNκμκ)2(x,y)|𝑑κ).{\mathbb{E}}\left\lparen F_{N}(\underline{x}_{N}^{t},\mu^{t})-F_{N}(\underline{x}_{N}^{0},\mu^{0})\right\rparen\leq 2\sigma{\mathbb{E}}\left\lparen\int_{0}^{t}\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}\Delta{\mathsf{g}}(x-y)d(\mu_{N}^{\kappa}-\mu^{\kappa})^{\otimes 2}(x,y)d\kappa\right\rparen\\ +{\mathbb{E}}\left\lparen\int_{0}^{t}\left|\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}\left\lparen u^{\kappa}(x)-u^{\kappa}(y)\right\rparen\cdot\nabla{\mathsf{g}}(x-y)d(\mu_{N}^{\kappa}-\mu^{\kappa})^{\otimes 2}(x,y)\right|d\kappa\right\rparen.
Remark 6.4.

By Proposition 5.1, Proposition 6.3 also implies that there is a constant C>0C>0 such that

(6.17) 𝔼(|FN(x¯Nt,μt)||FN(x¯N0,μ0)|)C(ηs(1+|logη|𝟏s=0)N+η2+μLηds(1+|logη|𝟏s=0))+2σ𝔼(0t(d)2Δ𝗀(xy)d(μNκμκ)2(x,y)𝑑κ)+𝔼(0t|(d)2(uκ(x)uκ(y))𝗀(xy)d(μNκμκ)2(x,y)|𝑑κ){\mathbb{E}}\left\lparen\left|F_{N}(\underline{x}_{N}^{t},\mu^{t})\right|-\left|F_{N}(\underline{x}_{N}^{0},\mu^{0})\right|\right\rparen\leq C\left\lparen\frac{\eta^{-s}(1+|\log\eta|\mathbf{1}_{s=0})}{N}+\eta^{2}+\|\mu\|_{L^{\infty}}\eta^{d-s}(1+|\log\eta|\mathbf{1}_{s=0})\right\rparen\\ +2\sigma{\mathbb{E}}\left\lparen\int_{0}^{t}\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}\Delta{\mathsf{g}}(x-y)d(\mu_{N}^{\kappa}-\mu^{\kappa})^{\otimes 2}(x,y)d\kappa\right\rparen\\ +{\mathbb{E}}\left\lparen\int_{0}^{t}\left|\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}\left\lparen u^{\kappa}(x)-u^{\kappa}(y)\right\rparen\cdot\nabla{\mathsf{g}}(x-y)d(\mu_{N}^{\kappa}-\mu^{\kappa})^{\otimes 2}(x,y)\right|d\kappa\right\rparen

for any choice of 0<η<min{12,r02}0<\eta<\min\{\frac{1}{2},\frac{r_{0}}{2}\}. Similarly, using Proposition 5.8 if 𝗀{\mathsf{g}} is globally superharmonic, Proposition 6.3 also implies that for any p>ds+2\infty\geq p>\frac{d}{s+2}, there is a Cp>0C_{p}>0 such that

(6.18) 𝔼(|FN(x¯Nt,μt)||FN(x¯N0,μ0)|)CpμLγs,pηλs,p(1+(|logη|+|logμL|)𝟏s=0)+Cηs(1+|logη|𝟏s=0)N+2σ𝔼(0t(d)2Δ𝗀(xy)d(μNκμκ)2(x,y)𝑑κ)+𝔼(0t|(d)2(uκ(x)uκ(y))𝗀(xy)d(μNκμκ)2(x,y)|𝑑κ){\mathbb{E}}\left\lparen\left|F_{N}(\underline{x}_{N}^{t},\mu^{t})\right|-\left|F_{N}(\underline{x}_{N}^{0},\mu^{0})\right|\right\rparen\leq C_{p}\|\mu\|_{L^{\infty}}^{\gamma_{s,p}}\eta^{\lambda_{s,p}}\left\lparen 1+\left\lparen|\log\eta|+|\log\|\mu\|_{L^{\infty}}|\right\rparen\mathbf{1}_{s=0}\right\rparen\\ +\frac{C\eta^{-s}(1+|\log\eta|\mathbf{1}_{s=0})}{N}+2\sigma{\mathbb{E}}\left\lparen\int_{0}^{t}\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}\Delta{\mathsf{g}}(x-y)d(\mu_{N}^{\kappa}-\mu^{\kappa})^{\otimes 2}(x,y)d\kappa\right\rparen\\ +{\mathbb{E}}\left\lparen\int_{0}^{t}\left|\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}\left\lparen u^{\kappa}(x)-u^{\kappa}(y)\right\rparen\cdot\nabla{\mathsf{g}}(x-y)d(\mu_{N}^{\kappa}-\mu^{\kappa})^{\otimes 2}(x,y)\right|d\kappa\right\rparen

for any 0<η<2dpd+2pd(p1)μL1d0<\eta<2^{-\frac{dp-d+2p}{d(p-1)}}\|\mu\|_{L^{\infty}}^{-\frac{1}{d}}.

Proof of Proposition 6.3.

Fix t>0t>0. By mollifying the initial datum and using the continuous dependence in Proposition 3.1, we assume without loss of generality that μ\mu is CC^{\infty}. Since

(6.19) 22σNi=1N0()P.V.d{xi,εκ}𝗀(ε)(xi,εκy)d(μN,εκμκ)(y)𝑑Wiκ\begin{split}&\frac{2\sqrt{2\sigma}}{N}\sum_{i=1}^{N}\int_{0}^{(\cdot)}\PV\int_{{\mathbb{R}}^{d}\setminus\{x_{i,\varepsilon}^{\kappa}\}}\nabla{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}^{\kappa}-y)d(\mu_{N,\varepsilon}^{\kappa}-\mu^{\kappa})(y)\cdot dW_{i}^{\kappa}\end{split}

is a sum of square-integrable martingales with zero initial expectation, Doob’s optional sampling theorem implies that for every ε>0\varepsilon>0,

(6.20) 𝔼(22σNi=1N0τεtP.V.d{xi,εκ}𝗀(ε)(xi,εκy)d(μN,εκμκ)(y)𝑑Wiκ)=0.{\mathbb{E}}\left\lparen\frac{2\sqrt{2\sigma}}{N}\sum_{i=1}^{N}\int_{0}^{\tau_{\varepsilon}\wedge t}\PV\int_{{\mathbb{R}}^{d}\setminus\{x_{i,\varepsilon}^{\kappa}\}}\nabla{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}^{\kappa}-y)d(\mu_{N,\varepsilon}^{\kappa}-\mu^{\kappa})(y)\cdot dW_{i}^{\kappa}\right\rparen=0.

Next, consider the expression

(6.21) (d)2Δ(𝗀(ε)𝗀)(xy)d(μN,εκμκ)2(x,y).\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}\Delta({\mathsf{g}}_{(\varepsilon)}-{\mathsf{g}})(x-y)d(\mu_{N,\varepsilon}^{\kappa}-\mu^{\kappa})^{\otimes 2}(x,y).

Observe that by definition (4.2) of 𝗀(ε){\mathsf{g}}_{(\varepsilon)},

(6.22) (𝗀(ε)𝗀)Δμκ(x)=d𝗀(xy)χε(xy)Δμκ(y)𝑑y.({\mathsf{g}}_{(\varepsilon)}-{\mathsf{g}})\ast\Delta\mu^{\kappa}(x)=-\int_{{\mathbb{R}}^{d}}{\mathsf{g}}(x-y)\chi_{\varepsilon}(x-y)\Delta\mu^{\kappa}(y)dy.

Integrating by parts twice to move the derivatives off μκ\mu^{\kappa} and then applying Cauchy-Schwarz and using iv for 𝗀{\mathsf{g}}, we find

(6.23) |d𝗀(xy)χε(xy)Δμκ(y)𝑑y|\displaystyle\left|\int_{{\mathbb{R}}^{d}}{\mathsf{g}}(x-y)\chi_{\varepsilon}(x-y)\Delta\mu^{\kappa}(y)dy\right| μκLε2+s(|xy|2ε𝑑y)μ0Lεd2s,\displaystyle\lesssim\frac{\|\mu^{\kappa}\|_{L^{\infty}}}{\varepsilon^{2+s}}\left\lparen\int_{|x-y|\leq 2\varepsilon}dy\right\rparen\lesssim\|\mu^{0}\|_{L^{\infty}}\varepsilon^{d-2-s},

where in the ultimate inequality we use the nonincreasing property of LpL^{p} norms. Thus,

(6.24) |d(𝗀(ε)𝗀)Δμκ(x)d(μN,εκμκ)(x)|μ0Lεd2s.\left|\int_{{\mathbb{R}}^{d}}({\mathsf{g}}_{(\varepsilon)}-{\mathsf{g}})\ast\Delta\mu^{\kappa}(x)d(\mu_{N,\varepsilon}^{\kappa}-\mu^{\kappa})(x)\right|\lesssim\|\mu^{0}\|_{L^{\infty}}\varepsilon^{d-2-s}.

Since for every 0κτε0\leq\kappa\leq\tau_{\varepsilon},

(6.25) 1ijNΔ𝗀(ε)(xi,εκxj,εκ)=1ijNΔ𝗀(xiκxjκ),\sum_{1\leq i\neq j\leq N}\Delta{\mathsf{g}}_{(\varepsilon)}(x_{i,\varepsilon}^{\kappa}-x_{j,\varepsilon}^{\kappa})=\sum_{1\leq i\neq j\leq N}\Delta{\mathsf{g}}(x_{i}^{\kappa}-x_{j}^{\kappa}),

using limε0τε=\lim_{\varepsilon\to 0}\tau_{\varepsilon}=\infty a.s., it follows that with probability one, for all 0κt0\leq\kappa\leq t,

(6.26) limε0+|(d)2(Δ𝗀(ε)(xy)d(μN,εκμκ)2(x,y)Δ𝗀(xy)d(μNκμκ)2(x,y))|=0.\lim_{\varepsilon\rightarrow 0^{+}}\left|\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}\left\lparen\Delta{\mathsf{g}}_{(\varepsilon)}(x-y)d(\mu_{N,\varepsilon}^{\kappa}-\mu^{\kappa})^{\otimes 2}(x,y)-\Delta{\mathsf{g}}(x-y)d(\mu_{N}^{\kappa}-\mu^{\kappa})^{\otimes 2}(x,y)\right\rparen\right|=0.

So by dominated convergence,

(6.27) limε0+𝔼(0tτε(d)2Δ𝗀(ε)(xy)d(μN,εκμκ)2(x,y)𝑑κ)=𝔼(0t(d)2Δ𝗀(xy)d(μNκμκ)2(x,y)𝑑κ).\begin{split}\lim_{\varepsilon\rightarrow 0^{+}}{\mathbb{E}}\left\lparen\int_{0}^{t\wedge\tau_{\varepsilon}}\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}\Delta{\mathsf{g}}_{(\varepsilon)}(x-y)d(\mu_{N,\varepsilon}^{\kappa}-\mu^{\kappa})^{\otimes 2}(x,y)d\kappa\right\rparen\\ ={\mathbb{E}}\left\lparen\int_{0}^{t}\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}\Delta{\mathsf{g}}(x-y)d(\mu_{N}^{\kappa}-\mu^{\kappa})^{\otimes 2}(x,y)d\kappa\right\rparen.\end{split}

Next, consider the expression

(6.28) 0t(d)2(uεκ(x)uεκ(y))𝗀(ε)(xy)d(μN,εκμκ)2(x,y)𝑑κ.\int_{0}^{t}\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}\left\lparen u_{\varepsilon}^{\kappa}(x)-u_{\varepsilon}^{\kappa}(y)\right\rparen\cdot\nabla{\mathsf{g}}_{(\varepsilon)}(x-y)d(\mu_{N,\varepsilon}^{\kappa}-\mu^{\kappa})^{\otimes 2}(x,y)d\kappa.

We want to show that

(6.29) limε0𝔼(0tτε|(d)2(uεκ(x)uεκ(y))𝗀(ε)(xy)d(μN,εκμκ)2(x,y)(d)2(uκ(x)uκ(y))𝗀(xy)d(μNκμκ)2(x,y)|dκ)=0.\lim_{\varepsilon\rightarrow 0}{\mathbb{E}}\Bigg{(}\int_{0}^{t\wedge\tau_{\varepsilon}}\left|\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}\left\lparen u_{\varepsilon}^{\kappa}(x)-u_{\varepsilon}^{\kappa}(y)\right\rparen\cdot\nabla{\mathsf{g}}_{(\varepsilon)}(x-y)d(\mu_{N,\varepsilon}^{\kappa}-\mu^{\kappa})^{\otimes 2}(x,y)\right.\\ -\left.\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}\left\lparen u^{\kappa}(x)-u^{\kappa}(y)\right\rparen\cdot\nabla{\mathsf{g}}(x-y)d(\mu_{N}^{\kappa}-\mu^{\kappa})^{\otimes 2}(x,y)\right|d\kappa\Bigg{)}=0.

We break up the demonstration of (6.29) into three parts.

• Almost surely, we have that for all 0κτε0\leq\kappa\leq\tau_{\varepsilon}

(6.30) (d)2(uεκ(x)uεκ(y))𝗀(ε)(xy)d(μN,εκ)2(x,y)=1N21ijN(uεκ(xiκ)uεκ(xjκ))𝗀(xiκxjκ).\begin{split}&\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}\left\lparen u_{\varepsilon}^{\kappa}(x)-u_{\varepsilon}^{\kappa}(y)\right\rparen\cdot\nabla{\mathsf{g}}_{(\varepsilon)}(x-y)d(\mu_{N,\varepsilon}^{\kappa})^{\otimes 2}(x,y)\\ &=\frac{1}{N^{2}}\sum_{1\leq i\neq j\leq N}\left\lparen u_{\varepsilon}^{\kappa}(x_{i}^{\kappa})-u_{\varepsilon}^{\kappa}(x_{j}^{\kappa})\right\rparen\cdot\nabla{\mathsf{g}}(x_{i}^{\kappa}-x_{j}^{\kappa}).\end{split}

Write

(6.31) uuε=𝕄(𝗀𝗀(ε))μ.u-u_{\varepsilon}={\mathbb{M}}\nabla({\mathsf{g}}-{\mathsf{g}}_{(\varepsilon)})\ast\mu.

We see from the same reasoning as in the estimate (6.23) that

(6.32) 2(𝗀𝗀(ε))μκLμκLεds2μ0Lεds2.\|\nabla^{\otimes 2}({\mathsf{g}}-{\mathsf{g}}_{(\varepsilon)})\ast\mu^{\kappa}\|_{L^{\infty}}\lesssim\|\mu^{\kappa}\|_{L^{\infty}}\varepsilon^{d-s-2}\leq\|\mu^{0}\|_{L^{\infty}}\varepsilon^{d-s-2}.

Hence by the mean-value theorem and using assumption v for 𝗀{\mathsf{g}},

|(((𝗀(ε)𝗀)μκ)(xiκ)((𝗀(ε)𝗀)μκ)(xjκ))𝗀(xiκxjκ)|\displaystyle\left|\left\lparen(\nabla({\mathsf{g}}_{(\varepsilon)}-{\mathsf{g}})\ast\mu^{\kappa})(x_{i}^{\kappa})-(\nabla({\mathsf{g}}_{(\varepsilon)}-{\mathsf{g}})\ast\mu^{\kappa})(x_{j}^{\kappa})\right\rparen\cdot\nabla{\mathsf{g}}(x_{i}^{\kappa}-x_{j}^{\kappa})\right|
μ0Lεds2|xiκxjκ||𝗀(xiκxjκ)|\displaystyle\quad\lesssim\|\mu^{0}\|_{L^{\infty}}\varepsilon^{d-s-2}|x_{i}^{\kappa}-x_{j}^{\kappa}||\nabla{\mathsf{g}}(x_{i}^{\kappa}-x_{j}^{\kappa})|
(6.33) {μ0Lεd2,s=0μ0Lεd2s𝗀(xiκxjκ),0<s<d2,\displaystyle\quad\lesssim\begin{cases}\|\mu^{0}\|_{L^{\infty}}\varepsilon^{d-2},&{s=0}\\ \|\mu^{0}\|_{L^{\infty}}\varepsilon^{d-2-s}{\mathsf{g}}(x_{i}^{\kappa}-x_{j}^{\kappa}),&0<s<d-2,\end{cases}

where to obtain the ultimate line in the case 0<s<d20<s<d-2 we assume |xiκxjκ|<r0|x_{i}^{\kappa}-x_{j}^{\kappa}|<r_{0}. For the case s=0s=0, the preceding estimate suffices. For the case 0<s<d20<s<d-2, we need to deal with the factor 𝗀(xixj){\mathsf{g}}(x_{i}-x_{j}). To this end, we set HN(x¯N)1ijN𝗀(xixj)H_{N}(\underline{x}_{N})\coloneqq\sum_{1\leq i\neq j\leq N}{\mathsf{g}}(x_{i}-x_{j}). Since 𝗀{\mathsf{g}} is positive inside the ball B(0,r0)B(0,r_{0}) and |𝗀|C|{\mathsf{g}}|\leq C outside B(0,r0)B(0,r_{0}) by assumptions iv and v, we find that

𝔼(0tτεHN(x¯Nκ)𝑑κ)\displaystyle{\mathbb{E}}\left\lparen\int_{0}^{t\wedge\tau_{\varepsilon}}H_{N}(\underline{x}_{N}^{\kappa})d\kappa\right\rparen =𝔼(0tτεHN,ε(x¯N,εκ)𝑑κ)\displaystyle={\mathbb{E}}\left\lparen\int_{0}^{t\wedge\tau_{\varepsilon}}H_{N,\varepsilon}(\underline{x}_{N,\varepsilon}^{\kappa})d\kappa\right\rparen
𝔼(0t(HN,ε(x¯N,εκ)+CN2)𝑑κ)\displaystyle\quad\leq{\mathbb{E}}\left\lparen\int_{0}^{t}\left\lparen H_{N,\varepsilon}(\underline{x}_{N,\varepsilon}^{\kappa})+CN^{2}\right\rparen d\kappa\right\rparen
=0t𝔼(HN,ε(x¯N,εκ)+CN2)𝑑κ\displaystyle\quad=\int_{0}^{t}{\mathbb{E}}\left\lparen H_{N,\varepsilon}(\underline{x}_{N,\varepsilon}^{\kappa})+CN^{2}\right\rparen d\kappa
(6.34) t(𝔼(HN(x¯N0))+CN2),\displaystyle\quad\leq t\left\lparen{\mathbb{E}}(H_{N}(\underline{x}_{N}^{0}))+CN^{2}\right\rparen,

where the ultimate line follows from the proof of (4.24), which is valid for any tt, and the fact that we may always assume ε<minij|xi0xj0|\varepsilon<\min_{i\neq j}|x_{i}^{0}-x_{j}^{0}|. It now follows that

(6.35) 𝔼(0tτε|(d)2((uεκuκ)(x)(uεκuκ)(y))𝗀(ε)(xy)d(μN,εκ)2(x,y)|𝑑κ){\mathbb{E}}\left\lparen\int_{0}^{t\wedge\tau_{\varepsilon}}\left|\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}\left\lparen(u_{\varepsilon}^{\kappa}-u^{\kappa})(x)-(u_{\varepsilon}^{\kappa}-u^{\kappa})(y)\right\rparen\cdot\nabla{\mathsf{g}}_{(\varepsilon)}(x-y)d(\mu_{N,\varepsilon}^{\kappa})^{\otimes 2}(x,y)\right|d\kappa\right\rparen

vanishes as ε0+\varepsilon\rightarrow 0^{+}.

• Almost surely, for 0κτε0\leq\kappa\leq\tau_{\varepsilon},

(6.36) (d)2(uεκ(x)uεκ(y))𝗀(ε)(xy)𝑑μN,εκ(x)𝑑μκ(y)=1Ni=1N(𝕄𝗀(ε)μκ)(xiκ)(𝗀(ε)μκ)(xiκ)1Ni=1N(𝗀(ε)(div(μκuεκ)))(xiκ).\begin{split}&\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}\left\lparen u_{\varepsilon}^{\kappa}(x)-u_{\varepsilon}^{\kappa}(y)\right\rparen\cdot\nabla{\mathsf{g}}_{(\varepsilon)}(x-y)d\mu_{N,\varepsilon}^{\kappa}(x)d\mu^{\kappa}(y)\\ &=\frac{1}{N}\sum_{i=1}^{N}({\mathbb{M}}\nabla{\mathsf{g}}_{(\varepsilon)}\ast\mu^{\kappa})(x_{i}^{\kappa})\cdot(\nabla{\mathsf{g}}_{(\varepsilon)}\ast\mu^{\kappa})(x_{i}^{\kappa})-\frac{1}{N}\sum_{i=1}^{N}\left\lparen{\mathsf{g}}_{(\varepsilon)}\ast\left\lparen\operatorname{\mathrm{div}}(\mu^{\kappa}u_{\varepsilon}^{\kappa})\right\rparen\right\rparen(x_{i}^{\kappa}).\end{split}

By Remark 2.5, the same reasoning as (6.23) and the nonincreasing property of LpL^{p} norms, we have

|(𝕄𝗀(ε)μκ)(𝗀(ε)μκ)(𝕄𝗀μκ)(𝗀μκ)|\displaystyle\left|({\mathbb{M}}\nabla{\mathsf{g}}_{(\varepsilon)}\ast\mu^{\kappa})\cdot(\nabla{\mathsf{g}}_{(\varepsilon)}\ast\mu^{\kappa})-({\mathbb{M}}\nabla{\mathsf{g}}\ast\mu^{\kappa})\cdot(\nabla{\mathsf{g}}\ast\mu^{\kappa})\right|
(𝗀(ε)𝗀)μκL(𝗀(ε)μκL+𝗀μκL)\displaystyle\quad\leq\|\nabla({\mathsf{g}}_{(\varepsilon)}-{\mathsf{g}})\ast\mu^{\kappa}\|_{L^{\infty}}(\|\nabla{\mathsf{g}}_{(\varepsilon)}\ast\mu^{\kappa}\|_{L^{\infty}}+\|\nabla{\mathsf{g}}\ast\mu^{\kappa}\|_{L^{\infty}})
(6.37) εds1μ0Ld+s+1d.\displaystyle\quad\lesssim\varepsilon^{d-s-1}\|\mu^{0}\|_{L^{\infty}}^{\frac{d+s+1}{d}}.

Similarly,

|(𝗀(ε)div(μκuεκ))(xiκ)𝗀(div(μκuκ))(xiκ)|\displaystyle\left|\left\lparen{\mathsf{g}}_{(\varepsilon)}\ast\operatorname{\mathrm{div}}(\mu^{\kappa}u_{\varepsilon}^{\kappa})\right\rparen(x_{i}^{\kappa})-{\mathsf{g}}\ast\left\lparen\operatorname{\mathrm{div}}(\mu^{\kappa}u^{\kappa})\right\rparen(x_{i}^{\kappa})\right|
εds1μκuεκL+μκ(uεκuκ)L11s+1dμκ(uεκuκ)Ls+1d\displaystyle\quad\lesssim\varepsilon^{d-s-1}\|\mu^{\kappa}u_{\varepsilon}^{\kappa}\|_{L^{\infty}}+\|\mu^{\kappa}(u_{\varepsilon}^{\kappa}-u^{\kappa})\|_{L^{1}}^{1-\frac{s+1}{d}}\|\mu^{\kappa}(u_{\varepsilon}^{\kappa}-u^{\kappa})\|_{L^{\infty}}^{\frac{s+1}{d}}
(6.38) εds1μ0Ld+s+1d.\displaystyle\quad\lesssim\varepsilon^{d-s-1}\|\mu^{0}\|_{L^{\infty}}^{\frac{d+s+1}{d}}.

After a little bookkeeping, we find that

(6.39) 𝔼(0tτε|(d)2((uεκ(x)uεκ(y))𝗀(ε)(xy)dμN,εκ(x)dμκ(y)(d)2(uκ(x)uκ(y))𝗀(xy)dμNκ(x)dμκ(y)|dκ)tεds1μ0Ld+s+1d,{\mathbb{E}}\Bigg{(}\int_{0}^{t\wedge\tau_{\varepsilon}}\left|\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}\left\lparen(u_{\varepsilon}^{\kappa}(x)-u_{\varepsilon}^{\kappa}(y)\right\rparen\cdot\nabla{\mathsf{g}}_{(\varepsilon)}(x-y)d\mu_{N,\varepsilon}^{\kappa}(x)d\mu^{\kappa}(y)\right.\\ -\left.\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}\left\lparen u^{\kappa}(x)-u^{\kappa}(y)\right\rparen\cdot\nabla{\mathsf{g}}(x-y)d\mu_{N}^{\kappa}(x)d\mu^{\kappa}(y)\right|d\kappa\Bigg{)}\lesssim t\varepsilon^{d-s-1}\|\mu^{0}\|_{L^{\infty}}^{\frac{d+s+1}{d}},

which evidently vanishes as ε0+\varepsilon\rightarrow 0^{+}.

• Observe that

(6.40) (d)2(uεκ(x)uεκ(y))𝗀(ε)(xy)d(μκ)2(x,y)=2duεκ(x)(𝗀(ε)μκ)(x)dμκ(x)\begin{split}&\int_{({\mathbb{R}}^{d})^{2}\setminus\triangle}\left\lparen u_{\varepsilon}^{\kappa}(x)-u_{\varepsilon}^{\kappa}(y)\right\rparen\cdot\nabla{\mathsf{g}}_{(\varepsilon)}(x-y)d(\mu^{\kappa})^{\otimes 2}(x,y)=2\int_{{\mathbb{R}}^{d}}u_{\varepsilon}^{\kappa}(x)\cdot(\nabla{\mathsf{g}}_{(\varepsilon)}\ast\mu^{\kappa})(x)d\mu^{\kappa}(x)\end{split}

By Lemma 2.3 and the nonincreasing property of LpL^{p} norms, arguing similarly as above, we have

d|uεκ(x)(𝗀(ε)μκ)(x)uκ(x)(𝗀μκ)(x)|dμκ(x)\displaystyle\int_{{\mathbb{R}}^{d}}\left|u_{\varepsilon}^{\kappa}(x)\cdot(\nabla{\mathsf{g}}_{(\varepsilon)}\ast\mu^{\kappa})(x)-u^{\kappa}(x)\cdot(\nabla{\mathsf{g}}\ast\mu^{\kappa})(x)\right|d\mu^{\kappa}(x)
(𝗀𝗀(ε))μκL(𝗀(ε)μκL+𝗀μκL)\displaystyle\quad\leq\|\nabla({\mathsf{g}}-{\mathsf{g}}_{(\varepsilon)})\ast\mu^{\kappa}\|_{L^{\infty}}(\|\nabla{\mathsf{g}}_{(\varepsilon)}\ast\mu^{\kappa}\|_{L^{\infty}}+\|\nabla{\mathsf{g}}\ast\mu^{\kappa}\|_{L^{\infty}})
(6.41) εds1μ0Ld+s+1d.\displaystyle\quad\lesssim\varepsilon^{d-s-1}\|\mu^{0}\|_{L^{\infty}}^{\frac{d+s+1}{d}}.

Therefore,

(6.42) 𝔼(0tτεd|uεκ(x)(𝗀(ε)μκ)(x)uκ(x)(𝗀μκ)(x)|dμκ(x)dκ)tεds1μ0Ld+s+1d,{\mathbb{E}}\left\lparen\int_{0}^{t\wedge\tau_{\varepsilon}}\int_{{\mathbb{R}}^{d}}\left|u_{\varepsilon}^{\kappa}(x)\cdot(\nabla{\mathsf{g}}_{(\varepsilon)}\ast\mu^{\kappa})(x)-u^{\kappa}(x)\cdot(\nabla{\mathsf{g}}\ast\mu^{\kappa})(x)\right|d\mu^{\kappa}(x)d\kappa\right\rparen\lesssim t\varepsilon^{d-s-1}\|\mu^{0}\|_{L^{\infty}}^{\frac{d+s+1}{d}},

which evidently tends to zero as ε0+\varepsilon\rightarrow 0^{+}. With this last bit, the desired result (6.29) now follows.

Combining the above results, we see that we have shown the inequality

(6.43) limε0+𝔼(FN,ε(x¯N,ετεt,μτεt)FN(x¯N0,μ0))2σ𝔼(0tdΔ𝗀(xy)d(μNκμκ)2(x,y)dκ)+𝔼(0t|(2)2((uκ(x)uκ(y))𝗀(xy)d(μNκμκ)2(x,y)|dκ).\lim_{\varepsilon\rightarrow 0^{+}}{\mathbb{E}}\left\lparen F_{N,\varepsilon}(\underline{x}_{N,\varepsilon}^{\tau_{\varepsilon}\wedge t},\mu^{\tau_{\varepsilon}\wedge t})-F_{N}(\underline{x}_{N}^{0},\mu^{0})\right\rparen\leq 2\sigma{\mathbb{E}}\left\lparen\int_{0}^{t}\int_{{\mathbb{R}}^{d}}\Delta{\mathsf{g}}(x-y)d(\mu_{N}^{\kappa}-\mu^{\kappa})^{\otimes 2}(x,y)d\kappa\right\rparen\\ +{\mathbb{E}}\left\lparen\int_{0}^{t}\left|\int_{({\mathbb{R}}^{2})^{2}\setminus\triangle}\left\lparen(u^{\kappa}(x)-u^{\kappa}(y)\right\rparen\cdot\nabla{\mathsf{g}}(x-y)d(\mu_{N}^{\kappa}-\mu^{\kappa})^{\otimes 2}(x,y)\right|d\kappa\right\rparen.

Since FN,ε(x¯N,ε,μ)F_{N,\varepsilon}(\underline{x}_{N,\varepsilon},\mu) is bounded from below uniformly in the noise and time by virtue of Proposition 5.1, we can conclude the proof by applying Fatou’s lemma. ∎

7. Gronwall argument

We now have all the ingredients necessary to conclude our Gronwall argument for the modulated energy, thereby proving Theorems 1.1 and 1.2. We divide this section into two subsections. In the first subsection, we consider the case where the admissible potential 𝗀{\mathsf{g}} is only superharmonic in a neighborhood of the origin (i.e. r0<r_{0}<\infty in assumption iii). For such potentials, we obtain decay bounds for the modulated FN(x¯Nt,μt)F_{N}(\underline{x}_{N}^{t},\mu^{t}) as NN\rightarrow\infty which grow linearly in time. The conclusion is the proof of Theorem 1.1. In the second subsection, we consider admissible potentials which are superharmonic on d{\mathbb{R}}^{d} (i.e. r0=r_{0}=\infty in assumption iii). Under this stronger assumption, we can prove decay bounds for FN(x¯Nt,μt)F_{N}(\underline{x}_{N}^{t},\mu^{t}) which are uniform on the interval [0,)[0,\infty). The conclusion is the proof of Theorem 1.2.

7.1. Linear-in-time estimates

Applying 5.6 and Proposition 5.7 pointwise in time to the first and second terms, respectively, of the right-hand side of inequality (6.16) and using Remark 5.2 to control |FN(x¯Nt,μt)||F_{N}(\underline{x}_{N}^{t},\mu^{t})| in terms of FN(x¯Nt,μt)F_{N}(\underline{x}_{N}^{t},\mu^{t}), we find that

(7.1) 𝔼(|FN(x¯Nt,μt)|)|FN(x¯N0,μ0)|+C(1+μtL)N22+s(1+(logN)𝟏s=0)+Cσ0t(1+μκL)Nmin{2,ds2}min{s+4,d}dκ+C0t(uκL+||ds2uκL2dd2s)(FN(x¯Nκ,μκ)+CNs+3(s+2)(s+1)||s+1dμκL+C(1+μκL)N2(s+2)(s+1)(1+(logN)𝟏s=0))dκ,{\mathbb{E}}(|F_{N}(\underline{x}_{N}^{t},\mu^{t})|)\leq|F_{N}(\underline{x}_{N}^{0},\mu^{0})|+C(1+\|\mu^{t}\|_{L^{\infty}})N^{-\frac{2}{2+s}}\left\lparen 1+(\log N)\mathbf{1}_{s=0}\right\rparen\\ +C\sigma\int_{0}^{t}(1+\|\mu^{\kappa}\|_{L^{\infty}})N^{-\frac{\min\{2,d-s-2\}}{\min\{s+4,d\}}}d\kappa+C\int_{0}^{t}\Bigg{(}\|\nabla u^{\kappa}\|_{L^{\infty}}+\||\nabla|^{\frac{d-s}{2}}u^{\kappa}\|_{L^{\frac{2d}{d-2-s}}}\Bigg{)}\Bigg{(}F_{N}(\underline{x}_{N}^{\kappa},\mu^{\kappa})\\ +CN^{-\frac{s+3}{(s+2)(s+1)}}\||\nabla|^{s+1-d}\mu^{\kappa}\|_{L^{\infty}}+C(1+\|\mu^{\kappa}\|_{L^{\infty}})N^{-\frac{2}{(s+2)(s+1)}}\Big{(}1+(\log N)\mathbf{1}_{s=0}\Big{)}\Bigg{)}d\kappa,

where we have defined u𝕄𝗀μu\coloneqq{\mathbb{M}}\nabla{\mathsf{g}}\ast\mu. We remind the reader that the constant CC depends on r0r_{0} from assumption iii.

Using Remark 2.5 and the fact that μκL1=1\|\mu^{\kappa}\|_{L^{1}}=1, we see that

(7.2) uκLCμκLs+2d.\|\nabla u^{\kappa}\|_{L^{\infty}}\leq C\|\mu^{\kappa}\|_{L^{\infty}}^{\frac{s+2}{d}}.

Similarly, using the commutativity of Fourier multipliers together with the Hardy-Littlewood-Sobolev lemma, followed by Hölder’s inequality

||ds2uκL2dd2s=𝕄𝗀(||ds2μκ)L2dd2s\displaystyle\||\nabla|^{\frac{d-s}{2}}u^{\kappa}\|_{L^{\frac{2d}{d-2-s}}}=\|{\mathbb{M}}\nabla{\mathsf{g}}\ast(|\nabla|^{\frac{d-s}{2}}\mu^{\kappa})\|_{L^{\frac{2d}{d-2-s}}} Cds22(μκ)L2dd2s\displaystyle\leq C\|\mathcal{I}_{\frac{d-s-2}{2}}(\mu^{\kappa})\|_{L^{\frac{2d}{d-2-s}}}
CμκLdd2s\displaystyle\leq C\|\mu^{\kappa}\|_{L^{\frac{d}{d-2-s}}}
(7.3) CμκL2+sd.\displaystyle\leq C\|\mu^{\kappa}\|_{L^{\infty}}^{\frac{2+s}{d}}.

By another application of Lemma 2.3,

(7.4) ||s+1dμκLμκLs+1d.\||\nabla|^{s+1-d}\mu^{\kappa}\|_{L^{\infty}}\leq\|\mu^{\kappa}\|_{L^{\infty}}^{\frac{s+1}{d}}.

Applying the bounds (7.2), (7.3), (7.4) to the right-hand side of (7.1), then applying the Gronwall-Bellman lemma, we find that

(7.5) 𝔼(|FN(x¯Nt,μt)|)ANtexp(C0tμκLs+2ddκ),{\mathbb{E}}(|F_{N}(\underline{x}_{N}^{t},\mu^{t})|)\leq A_{N}^{t}\exp\left\lparen C\int_{0}^{t}\|\mu^{\kappa}\|_{L^{\infty}}^{\frac{s+2}{d}}d\kappa\right\rparen,

where the time-dependent prefactor ANtA_{N}^{t} is defined by

(7.6) ANt|FN(x¯N0,μ0)|+C(1+μtL)N22+s(1+(logN)𝟏s=0)+Cσ0t(1+μκL)Nmin{2,ds2}min{s+4,d}dκ+C0tμκLs+2d(μκLs+1dNs+3(s+2)(s+1)+C(1+μκL)N2(s+2)(s+1)(1+(logN)𝟏s=0))dκ.A_{N}^{t}\coloneqq|F_{N}(\underline{x}_{N}^{0},\mu^{0})|+C(1+\|\mu^{t}\|_{L^{\infty}})N^{-\frac{2}{2+s}}\left\lparen 1+(\log N)\mathbf{1}_{s=0}\right\rparen\\ +C\sigma\int_{0}^{t}(1+\|\mu^{\kappa}\|_{L^{\infty}})N^{-\frac{\min\{2,d-s-2\}}{\min\{s+4,d\}}}d\kappa+C\int_{0}^{t}\|\mu^{\kappa}\|_{L^{\infty}}^{\frac{s+2}{d}}\Bigg{(}\|\mu^{\kappa}\|_{L^{\infty}}^{\frac{s+1}{d}}N^{-\frac{s+3}{(s+2)(s+1)}}\\ +C(1+\|\mu^{\kappa}\|_{L^{\infty}})N^{-\frac{2}{(s+2)(s+1)}}\left\lparen 1+(\log N)\mathbf{1}_{s=0}\right\rparen\Bigg{)}d\kappa.

Without loss of generality, we may assume that t1t\geq 1. Split the interval [0,t][0,t] into [0,1][0,1] and [1,t][1,t]. On [0,1][0,1], we use the trivial bound μκLμ0L\|\mu^{\kappa}\|_{L^{\infty}}\leq\|\mu^{0}\|_{L^{\infty}}; and on [1,t][1,t], we use the bound μκLC(σκ)d2\|\mu^{\kappa}\|_{L^{\infty}}\leq C(\sigma\kappa)^{-\frac{d}{2}}, which comes from Proposition 3.8. It then follows that

(7.7) 0tμκLs+2ddκμ0Ls+2d+C1t(σκ)s+22dκμ0Ls+2d+C(2sσs+22𝟏s>0+(logt)σ𝟏s=0),\int_{0}^{t}\|\mu^{\kappa}\|_{L^{\infty}}^{\frac{s+2}{d}}d\kappa\leq\|\mu^{0}\|_{L^{\infty}}^{\frac{s+2}{d}}+C\int_{1}^{t}(\sigma\kappa)^{-\frac{s+2}{2}}d\kappa\leq\|\mu^{0}\|_{L^{\infty}}^{\frac{s+2}{d}}+C\left\lparen\frac{2}{s\sigma^{\frac{s+2}{2}}}\mathbf{1}_{s>0}+\frac{(\log t)}{\sigma}\mathbf{1}_{s=0}\right\rparen,
(7.8) 0tμκL2s+3ddκμ0L2s+3d+Cσ2s+32,\int_{0}^{t}\|\mu^{\kappa}\|_{L^{\infty}}^{\frac{2s+3}{d}}d\kappa\leq\|\mu^{0}\|_{L^{\infty}}^{\frac{2s+3}{d}}+\frac{C}{\sigma^{\frac{2s+3}{2}}},

and

(7.9) ANt|FN(x¯N0,μ0)|+C(1+μ0L)N22+s(1+(logN)𝟏s=0)+Cσt(1+μ0L)Nmin{2,ds2}min{s+4,d}+C(μ0L2s+3d+σ2s+32)Ns+3(s+2)(s+1)+Ct(1+μ0L)N2(s+2)(s+1)(1+(logN)𝟏s=0).A_{N}^{t}\leq|F_{N}(\underline{x}_{N}^{0},\mu^{0})|+C(1+\|\mu^{0}\|_{L^{\infty}})N^{-\frac{2}{2+s}}\left\lparen 1+(\log N)\mathbf{1}_{s=0}\right\rparen+C\sigma t(1+\|\mu^{0}\|_{L^{\infty}})N^{-\frac{\min\{2,d-s-2\}}{\min\{s+4,d\}}}\\ +C\left\lparen\|\mu^{0}\|_{L^{\infty}}^{\frac{2s+3}{d}}+\sigma^{-\frac{2s+3}{2}}\right\rparen N^{-\frac{s+3}{(s+2)(s+1)}}+Ct(1+\|\mu^{0}\|_{L^{\infty}})N^{-\frac{2}{(s+2)(s+1)}}\left\lparen 1+(\log N)\mathbf{1}_{s=0}\right\rparen.

Applying the preceding bounds and inserting into (7.5), we conclude

(7.10) 𝔼(|FN(x¯Nt,μt)|)exp(C(μ0Ls+2d+2sσs+22𝟏s>0+(logt1σ)𝟏s=0))(|FN(x¯N0,μ0)|+C(1+μ0L)N22+s(1+(logN)𝟏s=0)+Cσt(1+μ0L)Nmin{2,ds2}min{s+4,d}+C(μ0L2s+3d+σ2s+32)Ns+3(s+2)(s+1)+Ct(1+μ0L)N2(s+2)(s+1)(1+(logN)𝟏s=0)).{\mathbb{E}}(|F_{N}(\underline{x}_{N}^{t},\mu^{t})|)\leq\exp\left\lparen C\left\lparen\|\mu^{0}\|_{L^{\infty}}^{\frac{s+2}{d}}+\frac{2}{s\sigma^{\frac{s+2}{2}}}\mathbf{1}_{s>0}+(\log t^{\frac{1}{\sigma}})\mathbf{1}_{s=0}\right\rparen\right\rparen\Bigg{(}|F_{N}(\underline{x}_{N}^{0},\mu^{0})|\\ +C(1+\|\mu^{0}\|_{L^{\infty}})N^{-\frac{2}{2+s}}\left\lparen 1+(\log N)\mathbf{1}_{s=0}\right\rparen+C\sigma t(1+\|\mu^{0}\|_{L^{\infty}})N^{-\frac{\min\{2,d-s-2\}}{\min\{s+4,d\}}}\\ +C\left\lparen\|\mu^{0}\|_{L^{\infty}}^{\frac{2s+3}{d}}+\sigma^{-\frac{2s+3}{2}}\right\rparen N^{-\frac{s+3}{(s+2)(s+1)}}+Ct(1+\|\mu^{0}\|_{L^{\infty}})N^{-\frac{2}{(s+2)(s+1)}}\left\lparen 1+(\log N)\mathbf{1}_{s=0}\right\rparen\Bigg{)}.

Comparing (7.10) to (1.18), we see that we have proved Theorem 1.1.

7.2. Global-in-time estimates

We now assume that the potential 𝗀{\mathsf{g}} is globally superharmonic and show, using the results of Section 5.2, global-in-time bounds for the modulated energy FN(x¯Nt,μt)F_{N}(\underline{x}_{N}^{t},\mu^{t}) for the range 0<s<d20<s<d-2 and almost-global-in-time bounds if s=0s=0. This proves Theorem 1.2.

Applying 5.14 and Proposition 5.15 pointwise in time to the first and second terms, respectively, of the right-hand side of inequality (6.16) and using Remark 5.10 to control |FN(x¯Nt,μt)||F_{N}(\underline{x}_{N}^{t},\mu^{t})| in terms of FN(x¯Nt,μt)F_{N}(\underline{x}_{N}^{t},\mu^{t}), we find that

(7.11) 𝔼(|FN(x¯Nt,μt)|)|FN(x¯N0,μ0)|+CpμtLsdNλs,pλs,p+s(1+(|logμL|+logN)𝟏s=0)+Cσ0tμκLs+2d(CqNλs+2,qλs+2,q+s+2𝟏0sd4+Nds2d𝟏s>d4)dκ+C0tuκL||s+1dμκLNs+1+λs,p(s+λs,p)(1+s)dκ+C0t(uκL+||ds2uκL2dd2s)(FN(x¯Nκ,μκ)+Cp(1+μκLγs,p)Nλs,p(s+λs,p)(1+s)(1+(logN+|logμκL|)𝟏s=0))dκ{\mathbb{E}}\left\lparen|F_{N}(\underline{x}_{N}^{t},\mu^{t})|\right\rparen\leq|F_{N}(\underline{x}_{N}^{0},\mu^{0})|+C_{p}\|\mu^{t}\|_{L^{\infty}}^{\frac{s}{d}}N^{-\frac{\lambda_{s,p}}{\lambda_{s,p}+s}}\left\lparen 1+\left\lparen|\log\|\mu\|_{L^{\infty}}|+\log N\right\rparen\mathbf{1}_{s=0}\right\rparen\\ +C\sigma\int_{0}^{t}\|\mu^{\kappa}\|_{L^{\infty}}^{\frac{s+2}{d}}\left\lparen C_{q}N^{-\frac{\lambda_{s+2,q}}{\lambda_{s+2,q}+s+2}}\mathbf{1}_{0\leq s\leq d-4}+N^{-\frac{d-s-2}{d}}\mathbf{1}_{s>d-4}\right\rparen d\kappa\\ +C\int_{0}^{t}\|\nabla u^{\kappa}\|_{L^{\infty}}\||\nabla|^{s+1-d}\mu^{\kappa}\|_{L^{\infty}}N^{-\frac{s+1+\lambda_{s,p}}{(s+\lambda_{s,p})(1+s)}}d\kappa\\ +C\int_{0}^{t}\Bigg{(}\|\nabla u^{\kappa}\|_{L^{\infty}}+\||\nabla|^{\frac{d-s}{2}}u^{\kappa}\|_{L^{\frac{2d}{d-2-s}}}\Bigg{)}\Bigg{(}F_{N}(\underline{x}_{N}^{\kappa},\mu^{\kappa})\\ +C_{p}\left\lparen 1+\|\mu^{\kappa}\|_{L^{\infty}}^{\gamma_{s,p}}\right\rparen N^{-\frac{\lambda_{s,p}}{(s+\lambda_{s,p})(1+s)}}\left\lparen 1+\left\lparen\log N+|\log\|\mu^{\kappa}\|_{L^{\infty}}|\right\rparen\mathbf{1}_{s=0}\right\rparen\Bigg{)}d\kappa

for any choices p>ds+2\infty\geq p>\frac{d}{s+2} and q>ds+4\infty\geq q>\frac{d}{s+4}. The reader will recall the exponents γs,p,λs,p\gamma_{s,p},\lambda_{s,p} from (5.21). Implicit here is the assumption that N>(2dpd+p(p1)μκL)(s+1)(s+λs,p)dN>(2^{\frac{dp-d+p}{(p-1)}}\|\mu^{\kappa}\|_{L^{\infty}})^{\frac{(s+1)(s+\lambda_{s,p})}{d}} for every κ[0,t]\kappa\in[0,t], as required by Proposition 5.15. We can satisfy this constraint by assuming that N>(2dpd+p(p1)μ0L)(s+1)(s+λs,p)dN>(2^{\frac{dp-d+p}{(p-1)}}\|\mu^{0}\|_{L^{\infty}})^{\frac{(s+1)(s+\lambda_{s,p})}{d}}, since μκL\|\mu^{\kappa}\|_{L^{\infty}} is nonincreasing.

Applying the bounds (7.2), (7.3), (7.4) to the right-hand side of (7.11), then applying the Gronwall-Bellman lemma, we find that

(7.12) 𝔼(|FN(x¯Nt,μt)|)BNtexp(C0tμκLs+2ddκ),{\mathbb{E}}(|F_{N}(\underline{x}_{N}^{t},\mu^{t})|)\leq B_{N}^{t}\exp\left\lparen C\int_{0}^{t}\|\mu^{\kappa}\|_{L^{\infty}}^{\frac{s+2}{d}}d\kappa\right\rparen,

where the time-dependent prefactor BNtB_{N}^{t} is given by

(7.13) BNt|FN(x¯N0,μ0)|+CpμtLsdNλs,pλs,p+s(1+(|logμtL|+logN)𝟏s=0)+C0tμκL3+2sdNs+1+λs,p(s+λs,p)(1+s)dκ+Cσ0tμκLs+2d(CqNλs+2,qλs+2,q+s+2𝟏0sd4+Nds2d𝟏s>d4)dκ+Cp0tμκLs+2d(1+μκLγs,p)Nλs,p(s+λs,p)(1+s)(1+(logN+|logμκL|)𝟏s=0)dκ.B_{N}^{t}\coloneqq|F_{N}(\underline{x}_{N}^{0},\mu^{0})|+C_{p}\|\mu^{t}\|_{L^{\infty}}^{\frac{s}{d}}N^{-\frac{\lambda_{s,p}}{\lambda_{s,p}+s}}\left\lparen 1+\left\lparen|\log\|\mu^{t}\|_{L^{\infty}}|+\log N\right\rparen\mathbf{1}_{s=0}\right\rparen\\ +C\int_{0}^{t}\|\mu^{\kappa}\|_{L^{\infty}}^{\frac{3+2s}{d}}N^{-\frac{s+1+\lambda_{s,p}}{(s+\lambda_{s,p})(1+s)}}d\kappa+C\sigma\int_{0}^{t}\|\mu^{\kappa}\|_{L^{\infty}}^{\frac{s+2}{d}}\left\lparen C_{q}N^{-\frac{\lambda_{s+2,q}}{\lambda_{s+2,q}+s+2}}\mathbf{1}_{0\leq s\leq d-4}+N^{-\frac{d-s-2}{d}}\mathbf{1}_{s>d-4}\right\rparen d\kappa\\ +C_{p}\int_{0}^{t}\|\mu^{\kappa}\|_{L^{\infty}}^{\frac{s+2}{d}}\left\lparen 1+\|\mu^{\kappa}\|_{L^{\infty}}^{\gamma_{s,p}}\right\rparen N^{-\frac{\lambda_{s,p}}{(s+\lambda_{s,p})(1+s)}}\left\lparen 1+\left\lparen\log N+|\log\|\mu^{\kappa}\|_{L^{\infty}}|\right\rparen\mathbf{1}_{s=0}\right\rparen d\kappa.

Assuming t1t\geq 1 and splitting the interval [0,t][0,t] into [0,1],[1,t][0,1],[1,t] exactly as in the last subsection, we find that

(7.14) BNt|FN(x¯N0,μ0)|+C(μ0L3+2sd+σ3+2s2)Ns+1+λs,p(s+λs,p)(1+s)+Cpmin{μ0Lsd,(σt)s2}Nλs,pλs,p+s(1+(max{|logμ0L|,|log(σt)|}+logN)𝟏s=0)+Cσ(μ0Ls+2d+2sσs+22𝟏s>0+(logt1σ)𝟏s=0)(CqNλs+2,qλs+2,q+s+2𝟏0sd4+Nds2d𝟏s>d4)+Cp(1+μ0Lγs,p)(μ0Ls+2d+2sσs+22𝟏s>0+(logt1σ)𝟏s=0).B_{N}^{t}\leq|F_{N}(\underline{x}_{N}^{0},\mu^{0})|+C\left\lparen\|\mu^{0}\|_{L^{\infty}}^{\frac{3+2s}{d}}+\sigma^{-\frac{3+2s}{2}}\right\rparen N^{-\frac{s+1+\lambda_{s,p}}{(s+\lambda_{s,p})(1+s)}}\\ +C_{p}\min\{\|\mu^{0}\|_{L^{\infty}}^{\frac{s}{d}},(\sigma t)^{-\frac{s}{2}}\}N^{-\frac{\lambda_{s,p}}{\lambda_{s,p}+s}}\left\lparen 1+\left\lparen\max\{|\log\|\mu^{0}\|_{L^{\infty}}|,|\log(\sigma t)|\}+\log N\right\rparen\mathbf{1}_{s=0}\right\rparen\\ +C\sigma\left\lparen\|\mu^{0}\|_{L^{\infty}}^{\frac{s+2}{d}}+\frac{2}{s\sigma^{\frac{s+2}{2}}}\mathbf{1}_{s>0}+(\log t^{\frac{1}{\sigma}})\mathbf{1}_{s=0}\right\rparen\left\lparen C_{q}N^{-\frac{\lambda_{s+2,q}}{\lambda_{s+2,q}+s+2}}\mathbf{1}_{0\leq s\leq d-4}+N^{-\frac{d-s-2}{d}}\mathbf{1}_{s>d-4}\right\rparen\\ +C_{p}\left\lparen 1+\|\mu^{0}\|_{L^{\infty}}^{\gamma_{s,p}}\right\rparen\left\lparen\|\mu^{0}\|_{L^{\infty}}^{\frac{s+2}{d}}+\frac{2}{s\sigma^{\frac{s+2}{2}}}\mathbf{1}_{s>0}+(\log t^{\frac{1}{\sigma}})\mathbf{1}_{s=0}\right\rparen.

Applying this bound to the right-hand side of (7.12) and using (7.7) for the exponential factor, we conclude that

(7.15) 𝔼(|FN(x¯Nt,μt)|)exp(C(μ0Ls+2d+2sσs+22𝟏s>0+(logt1σ)𝟏s=0))(|FN(x¯N0,μ0)|+Cpmin{μ0Lsd,(σt)s2}Nλs,pλs,p+s(1+(max{|logμ0L|,|log(σt)|}+logN)𝟏s=0)+Cσ(μ0Ls+2d+2sσs+22𝟏s>0+(logt1σ)𝟏s=0)(CqNλs+2,qλs+2,q+s+2𝟏0sd4+Nds2d𝟏s>d4)+C(μ0L3+2sd+σ3+2s2)Ns+1+λs,p(s+λs,p)(1+s)+Cp(1+μ0Lγs,p)(μ0Ls+2d+2sσs+22𝟏s>0+(logt1σ)𝟏s=0)).{\mathbb{E}}\left\lparen|F_{N}(\underline{x}_{N}^{t},\mu^{t})|\right\rparen\leq\exp\left\lparen C\left\lparen\|\mu^{0}\|_{L^{\infty}}^{\frac{s+2}{d}}+\frac{2}{s\sigma^{\frac{s+2}{2}}}\mathbf{1}_{s>0}+(\log t^{\frac{1}{\sigma}})\mathbf{1}_{s=0}\right\rparen\right\rparen\Bigg{(}|F_{N}(\underline{x}_{N}^{0},\mu^{0})|\\ +C_{p}\min\{\|\mu^{0}\|_{L^{\infty}}^{\frac{s}{d}},(\sigma t)^{-\frac{s}{2}}\}N^{-\frac{\lambda_{s,p}}{\lambda_{s,p}+s}}\left\lparen 1+\left\lparen\max\{|\log\|\mu^{0}\|_{L^{\infty}}|,|\log(\sigma t)|\}+\log N\right\rparen\mathbf{1}_{s=0}\right\rparen\\ +C\sigma\left\lparen\|\mu^{0}\|_{L^{\infty}}^{\frac{s+2}{d}}+\frac{2}{s\sigma^{\frac{s+2}{2}}}\mathbf{1}_{s>0}+(\log t^{\frac{1}{\sigma}})\mathbf{1}_{s=0}\right\rparen\left\lparen C_{q}N^{-\frac{\lambda_{s+2,q}}{\lambda_{s+2,q}+s+2}}\mathbf{1}_{0\leq s\leq d-4}+N^{-\frac{d-s-2}{d}}\mathbf{1}_{s>d-4}\right\rparen\\ +C\left\lparen\|\mu^{0}\|_{L^{\infty}}^{\frac{3+2s}{d}}+\sigma^{-\frac{3+2s}{2}}\right\rparen N^{-\frac{s+1+\lambda_{s,p}}{(s+\lambda_{s,p})(1+s)}}+C_{p}\left\lparen 1+\|\mu^{0}\|_{L^{\infty}}^{\gamma_{s,p}}\right\rparen\left\lparen\|\mu^{0}\|_{L^{\infty}}^{\frac{s+2}{d}}+\frac{2}{s\sigma^{\frac{s+2}{2}}}\mathbf{1}_{s>0}+(\log t^{\frac{1}{\sigma}})\mathbf{1}_{s=0}\right\rparen\Bigg{)}.

References

  • [BCnC11] François Bolley, José A. Cañizo, and José A. Carrillo. Stochastic mean-field limit: non-Lipschitz forces and swarming. Math. Models Methods Appl. Sci., 21(11):2179–2210, 2011.
  • [BGM10] François Bolley, Arnaud Guillin, and Florent Malrieu. Trend to equilibrium and particle approximation for a weakly selfconsistent Vlasov-Fokker-Planck equation. M2AN Math. Model. Numer. Anal., 44(5):867–884, 2010.
  • [BJW19a] Didier Bresch, Pierre-Emmanuel Jabin, and Zhenfu Wang. Modulated free energy and mean field limit. Séminaire Laurent Schwartz–EDP et applications, pages 1–22, 2019.
  • [BJW19b] Didier Bresch, Pierre-Emmanuel Jabin, and Zhenfu Wang. On mean-field limits and quantitative estimates with a large class of singular kernels: application to the Patlak-Keller-Segel model. C. R. Math. Acad. Sci. Paris, 357(9):708–720, 2019.
  • [BJW20] Didier Bresch, Pierre-Emmanuel Jabin, and Zhenfu Wang. Mean-field limit and quantitative estimates with singular attractive kernels. arXiv preprint arXiv:2011.08022, 2020.
  • [BO19] Robert J. Berman and Magnus Önnheim. Propagation of chaos for a class of first order models with singular mean field interactions. SIAM J. Math. Anal., 51(1):159–196, 2019.
  • [Car91] Eric A. Carlen. Some integral identities and inequalities for entire functions and their application to the coherent state transform. J. Funct. Anal., 97(1):231–249, 1991.
  • [CB18] Lénaïc Chizat and Francis Bach. On the global convergence of gradient descent for over-parameterized models using optimal transport. In Proceedings of the 32nd International Conference on Neural Information Processing Systems, NIPS18, pages 3040–3050, Red Hook, NY, USA, 2018. Curran Associates Inc.
  • [CCH14] José Antonio Carrillo, Young-Pil Choi, and Maxime Hauray. The derivation of swarming models: mean-field limit and Wasserstein distances. In Collective dynamics from bacteria to crowds, volume 553 of CISM Courses and Lect., pages 1–46. Springer, Vienna, 2014.
  • [CFP12] José A. Carrillo, Lucas C. F. Ferreira, and Juliana C. Precioso. A mass-transportation approach to a one dimensional fluid mechanics model with nonlocal velocity. Adv. Math., 231(1):306–327, 2012.
  • [CGM08] P. Cattiaux, A. Guillin, and F. Malrieu. Probabilistic approach for granular media equations in the non-uniformly convex case. Probab. Theory Related Fields, 140(1-2):19–40, 2008.
  • [Cho73] Alexandre Joel Chorin. Numerical study of slightly viscous flow. J. Fluid Mech., 57(4):785–796, 1973.
  • [CL90] Eric A. Carlen and Michael Loss. Extremals of functionals with competing symmetries. J. Funct. Anal., 88(2):437–456, 1990.
  • [CL95] Eric A. Carlen and Michael Loss. Optimal smoothing and decay estimates for viscously damped conservation laws, with applications to the 22-D Navier-Stokes equation. Duke Math. J., 81(1):135 – 157, 1995.
  • [Dav87] E. B. Davies. Explicit constants for Gaussian upper bounds on heat kernels. Amer. J. Math., 109(2):319–333, 1987.
  • [DEGZ20] Alain Durmus, Andreas Eberle, Arnaud Guillin, and Raphael Zimmer. An elementary approach to uniform in time propagation of chaos. Proc. Amer. Math. Soc., 148(12):5387–5398, 2020.
  • [DT21] François Delarue and Alvin Tse. Uniform in time weak propagation of chaos on the torus. arXiv preprint arXiv:2104.14973, 2021.
  • [Due16] M. Duerinckx. Mean-field limits for some Riesz interaction gradient flows. SIAM Journal on Mathematical Analysis, 48(3):2269–2300, 2016.
  • [FHM14] Nicolas Fournier, Maxime Hauray, and Stéphane Mischler. Propagation of chaos for the 2D viscous vortex model. J. Eur. Math. Soc. (JEMS), 16(7):1423–1466, 2014.
  • [FJ17] Nicolas Fournier and Benjamin Jourdain. Stochastic particle approximation of the Keller-Segel equation and two-dimensional generalization of Bessel processes. Ann. Appl. Probab., 27(5):2807–2861, 2017.
  • [GBM21] Arnaud Guillin, Pierre Le Bris, and Pierre Monmarché. Uniform in time propagation of chaos for the 2d vortex model and other singular stochastic systems. arXiv preprint arXiv:2108.08675, 2021.
  • [Gol16] François Golse. On the dynamics of large particle systems in the mean field limit. In Macroscopic and large scale phenomena: coarse graining, mean field limits and ergodicity, volume 3 of Lect. Notes Appl. Math. Mech., pages 1–144. Springer, 2016.
  • [GQn15] David Godinho and Cristobal Quiñinao. Propagation of chaos for a subcritical Keller-Segel model. Annales de l’I.H.P. Probabilités et statistiques, 51(3):965–992, 2015.
  • [Gra14a] Loukas Grafakos. Classical Fourier Analysis. Number 249 in Graduate Texts in Mathematics. Springer, third edition, 2014.
  • [Gra14b] Loukas Grafakos. Modern Fourier Analysis. Number 250 in Graduate Texts in Mathematics. Springer, third edition, 2014.
  • [Gro75] Leonard Gross. Logarithmic Sobolev inequalities. Amer. J. Math., 97(4):1061–1083, 1975.
  • [GW05] Thierry Gallay and C. Eugene Wayne. Global stability of vortex solutions of the two-dimensional Navier-Stokes equation. Comm. Math. Phys., 255(1):97–129, 2005.
  • [Hau09] Maxime Hauray. Wasserstein distances for vortices approximation of Euler-type equations. Math. Models Methods Appl. Sci., 19(8):1357–1384, 2009.
  • [HK02] Rainer Hegselmann and Ulrich Krause. Opinion Dynamics and Bounded Confidence Models, Analysis and Simulation. Journal of Artificial Societies and Social Simulation, 5(3):1–2, 2002.
  • [HM14] Maxime Hauray and Stéphane Mischler. On Kac’s chaos and related problems. J. Funct. Anal., 266(10):6055–6157, 2014.
  • [Hol16] Thomas Holding. Propagation of chaos for Hölder continuous interaction kernels via Glivenko-Cantelli. arXiv preprint arXiv:1608.02877, 2016.
  • [Jab14] Pierre-Emmanuel Jabin. A review of the mean field limits for Vlasov equations. Kinet. Relat. Models, 7(4):661–711, 2014.
  • [JW16] Pierre-Emmanuel Jabin and Zhenfu Wang. Mean field limit and propagation of chaos for Vlasov systems with bounded forces. J. Funct. Anal., 271(12):3588–3627, 2016.
  • [JW17] Pierre-Emmanuel Jabin and Zhenfu Wang. Mean field limit for stochastic particle systems. In Act. Part. Vol. 1. Advances theory, Model. Appl., Model. Simul. Sci. Eng. Technol., pages 379–402. Birkhäuser/Springer, Cham, 2017.
  • [JW18] Pierre-Emmanuel Jabin and Zhenfu Wang. Quantitative estimates of propagation of chaos for stochastic systems with W1,W^{-1,\infty} kernels. Invent. Math., 214(1):523–591, 2018.
  • [Kra00] U. Krause. A discrete nonlinear and non-autonomous model of consensus formation. Communications in Difference Equations, 2000.
  • [KS91] Ioannis Karatzas and Steven E. Shreve. Brownian motion and stochastic calculus, volume 113 of Graduate Texts in Mathematics. Springer-Verlag, New York, second edition, 1991.
  • [Lac21] Daniel Lacker. Hierarchies, entropy, and quantitative propagation of chaos for mean field diffusions. arXiv preprint arXiv:2105.02983, 2021.
  • [LLY19] Lei Li, Jian-Guo Liu, and Pu Yu. On the mean field limit for Brownian particles with Coulomb interaction in 3D. J. Math. Phys., 60(11):111501, 34, 2019.
  • [LY16] Jian-Guo Liu and Rong Yang. Propagation of chaos for large Brownian particle system with Coulomb interaction. Res. Math. Sci., 3:Paper No. 40, 33, 2016.
  • [M9́6] Sylvie Méléard. Asymptotic behaviour of some interacting particle systems; McKean-Vlasov and Boltzmann models. In Probabilistic models for nonlinear partial differential equations (Montecatini Terme, 1995), volume 1627 of Lecture Notes in Math., pages 42–95. Springer, Berlin, 1996.
  • [Mae08] Yasunori Maekawa. A lower bound for fundamental solutions of the heat convection equations. Arch. Ration. Mech. Anal., 189(1):45–58, 2008.
  • [Mal03] Florent Malrieu. Convergence to equilibrium for granular media equations and their Euler schemes. Ann. Appl. Probab., 13(2):540–560, 2003.
  • [McK67] H. P. McKean, Jr. Propagation of chaos for a class of non-linear parabolic equations. In Stochastic Differential Equations (Lecture Series in Differential Equations, Session 7, Catholic Univ., 1967), pages 41–57. Air Force Office Sci. Res., Arlington, Va., 1967.
  • [MMN18] Song Mei, Andrea Montanari, and Phan-Minh Nguyen. A mean field view of the landscape of two-layer neural networks. Proc. Natl. Acad. Sci. USA, 115(33):E7665–E7671, 2018.
  • [MMW15] Stéphane Mischler, Clément Mouhot, and Bernt Wennberg. A new approach to quantitative propagation of chaos for drift, diffusion and jump processes. Probab. Theory Related Fields, 161(1-2):1–59, 2015.
  • [MP12] Carlo Marchioro and Mario Pulvirenti. Mathematical theory of incompressible nonviscous fluids, volume 96. Springer Science & Business Media, 2012.
  • [MT11] Sebastien Motsch and Eitan Tadmor. A new model for self-organized dynamics and its flocking behavior. J. Stat. Phys., 144(5):923–947, 2011.
  • [Nas58] J. Nash. Continuity of solutions of parabolic and elliptic equations. Amer. J. Math., 80:931–954, 1958.
  • [NRS21] Quoc Hung Nguyen, Matthew Rosenzweig, and Sylvia Serfaty. Mean-field limits of Riesz-type singular flows with possible multiplicative transport noise. arXiv preprint arXiv:2107.02592, 2021.
  • [Ons49] L. Onsager. Statistical hydrodynamics. Nuovo Cimento (9), 6(Supplemento, 2 (Convegno Internazionale di Meccanica Statistica)):279–287, 1949.
  • [OS97] Hans G. Othmer and Angela Stevens. Aggregation, blowup, and collapse: the ABCs of taxis in reinforced random walks. SIAM J. Appl. Math., 57(4):1044–1081, 1997.
  • [Osa86] Hirofumi Osada. Propagation of chaos for the two-dimensional Navier-Stokes equation. Proc. Japan Acad. Ser. A Math. Sci., 62(1):8–11, 1986.
  • [Osa87a] Hirofumi Osada. Diffusion processes with generators of generalized divergence form. J. Math. Kyoto Univ., 27(4):597–619, 1987.
  • [Osa87b] Hirofumi Osada. Limit points of empirical distributions of vortices with small viscosity. In Hydrodynamic behavior and interacting particle systems (Minneapolis, Minn., 1986), volume 9 of IMA Vol. Math. Appl., pages 117–126. Springer, New York, 1987.
  • [Osa87c] Hirofumi Osada. Propagation of chaos for the two-dimensional Navier-Stokes equation. In Probabilistic methods in mathematical physics (Katata/Kyoto, 1985), pages 303–334. Academic Press, Boston, MA, 1987.
  • [Per07] Benoît Perthame. Transport equations in biology. Frontiers in Mathematics. Birkhäuser Verlag, Basel, 2007.
  • [Ros20a] Matthew Rosenzweig. Mean-field convergence of point vortices without regularity. arXiv preprint arXiv:2004.04140, 2020.
  • [Ros20b] Matthew Rosenzweig. The mean-field limit of stochastic point vortex systems with multiplicative noise. arXiv preprint arXiv:2011.12180, 2020.
  • [RS16] Nicolas Rougerie and Sylvia Serfaty. Higher-dimensional Coulomb gases and renormalized energy functionals. Comm. Pure Appl. Math., 69(3):519–605, 2016.
  • [RVE18] Grant M Rotskoff and Eric Vanden-Eijnden. Trainability and accuracy of neural networks: An interacting particle system approach. arXiv preprint arXiv:1805.00915, 2018.
  • [Ser17] Sylvia Serfaty. Mean field limits of the Gross-Pitaevskii and parabolic Ginzburg-Landau equations. J. Amer. Math. Soc., 30(3):713–768, 2017.
  • [Ser20] Sylvia Serfaty. Mean field limit for Coulomb-type flows. Duke Math. J., 169(15):2887–2935, 10 2020. Appendix with Mitia Duerinckx.
  • [SM93] Elias M. Stein and Timothy S. Murphy. Harmonic Analysis (PMS-43): Real-Variable Methods, Orthogonality, and Oscillatory Integrals. (PMS-43). Princeton University Press, 1993.
  • [Ste70] Elias M Stein. Singular Integrals and Differentiability Properties of Functions, volume 2. Princeton University Press, 1970.
  • [SV14] Sylvia Serfaty and Juan Luis Vázquez. A mean field equation as limit of nonlinear diffusions with fractional Laplacian operators. Calc. Var. Partial Differential Equations, 49(3-4):1091–1120, 2014.
  • [Szn91] Alain-Sol Sznitman. Topics in propagation of chaos. In École d’Été de Probabilités de Saint-Flour XIX—1989, volume 1464 of Lecture Notes in Math., pages 165–251. Springer, Berlin, 1991.
  • [TBL06] Chad M. Topaz, Andrea L. Bertozzi, and Mark A. Lewis. A nonlocal continuum model for biological aggregation. Bull. Math. Biol., 68(7):1601–1623, 2006.
  • [WZZ21] Zhenfu Wang, Xianliang Zhao, and Rongchan Zhu. Gaussian fluctuations for interacting particle systems with singular kernels. arXiv preprint arXiv:2105.13201, 2021.
  • [XWX11] Haoxiang Xia, Huili Wang, and Zhaoguo Xuan. Opinion dynamics: A multidisciplinary review and perspective on future research. Int. J. Knowl. Syst. Sci., 2(4):72–91, October 2011.