Boundary renormalisation of SPDEs

51
arXiv:2110.03656v1 [math.PR] 7 Oct 2021 Boundary renormalisation of SPDEs October 8, 2021 Máté Gerencsér 1 and Martin Hairer 2 1 Technische Universität Wien, Austria, Email: [email protected] 2 Imperial College London, UK, Email: [email protected] Abstract We consider the continuum parabolic Anderson model (PAM) and the dynamical Φ 4 equation on the 3-dimensional cube with boundary conditions. While the Dirichlet solution theories are relatively standard, the case of Neumann / Robin boundary conditions gives rise to a divergent boundary renormalisation. Furthermore for Φ 4 3 a ‘boundary triviality’ result is obtained: if one approximates the equation with Neumann boundary conditions and the usual bulk renormalisation, then the limiting process coincides with the one obtained using Dirichlet boundary conditions. Keywords: Boundary renormalisation, singular SPDE, regularity structures MSC classification: 60H15, 60L30 Contents 1 Introduction 2 1.1 The parabolic Anderson model ............................ 2 1.2 The dynamical Φ 4 3 equation .............................. 3 1.3 The linearisation step ................................. 4 2 Preliminaries 6 2.1 Hölder spaces ..................................... 6 2.2 Kernel bounds ..................................... 10 2.3 Convergence of the Robin kernels .......................... 14 3 Explicit boundary corrections 17 3.1 PAM .......................................... 17 3.2 Φ 4 3 - the quadratic term ................................ 20 3.3 Φ 4 3 - the cubic term .................................. 25 4 Regularity structures and models 29 4.1 General remarks .................................... 29 4.2 PAM .......................................... 34 4.3 Φ 4 3 ........................................... 37 5 Proofs of the main results 38 5.1 PAM .......................................... 40 5.2 Φ 4 3 ........................................... 43 A Abstract Schauder estimate – the Dirichlet case 46

Transcript of Boundary renormalisation of SPDEs

Page 1: Boundary renormalisation of SPDEs

arX

iv:2

110.

0365

6v1

[m

ath.

PR]

7 O

ct 2

021

Boundary renormalisation of SPDEs

October 8, 2021

Máté Gerencsér1 and Martin Hairer2

1 Technische Universität Wien, Austria, Email: [email protected] Imperial College London, UK, Email: [email protected]

Abstract

We consider the continuum parabolic Anderson model (PAM) and the dynamical Φ4

equation on the 3-dimensional cube with boundary conditions. While the Dirichlet solution

theories are relatively standard, the case of Neumann / Robin boundary conditions gives rise

to a divergent boundary renormalisation. Furthermore for Φ4

3a ‘boundary triviality’ result

is obtained: if one approximates the equation with Neumann boundary conditions and the

usual bulk renormalisation, then the limiting process coincides with the one obtained using

Dirichlet boundary conditions.

Keywords: Boundary renormalisation, singular SPDE, regularity structures

MSC classification: 60H15, 60L30

Contents

1 Introduction 2

1.1 The parabolic Anderson model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

1.2 The dynamical Φ4

3equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.3 The linearisation step . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

2 Preliminaries 6

2.1 Hölder spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

2.2 Kernel bounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

2.3 Convergence of the Robin kernels . . . . . . . . . . . . . . . . . . . . . . . . . . 14

3 Explicit boundary corrections 17

3.1 PAM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

3.2 Φ4

3- the quadratic term . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

3.3 Φ43 - the cubic term . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

4 Regularity structures and models 29

4.1 General remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

4.2 PAM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

4.3 Φ4

3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

5 Proofs of the main results 38

5.1 PAM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

5.2 Φ4

3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

A Abstract Schauder estimate – the Dirichlet case 46

Page 2: Boundary renormalisation of SPDEs

2 Introduction

1 Introduction

Most of the existing literature on singular SPDEs (and associated operators) considers

equations on flat domains without boundaries, like Td or Rd. There are also some recent

results where boundary conditions are considered and this raises analytical complications

but where the final statement is completely analogous to the periodic case. Examples are

[CvZ21], [Lab19], [GH19, Thms. 1.1, 1.5], [HP21, Thm. 1.1].

In some situations, however, the effect of the boundaries is more drastic. A notable

example is the open KPZ equation for which both in its derivation from exclusion pro-

cesses [CS18, Par18, GPS20] and its solution theory via regularity structures [GH19] an

approximation-dependent finite boundary correction term arises. A similar phenomenon

was observed in the context of a stochastic homogenisation problem in [HP21].

The goal of the present article is to extend the list of nontrivial boundary effects by

two well-known SPDEs endowed with appropriate boundary conditions. In contrast to

the previous examples, the boundary renormalisation in these examples does not remain

bounded but diverges logarithmically. A peculiar consequence is that if this divergence

comes with a positive sign (which turns out to be the case for the dynamical Φ43 model)

then removing the boundary renormalisation does not destroy the convergence of the Wong–

Zakai approximations but rather “trivialises” the boundary condition to a vanishing Dirichlet

one. More precisely, the sequence of smooth approximations with any fixed boundary

condition of Robin type (including Neumann) converges to the same limit as that with

Dirichlet boundary conditions, but it is possible to choose an ε-dependent Robin boundary

condition in such a way that the limit still exists but is different. This is a mechanism

somewhat analogous to that underlying the triviality result of [HRW12], although here the

limit is still nontrivial in the bulk.

1.1 The parabolic Anderson model

Consider the 3-dimensional continuum parabolic Anderson model formally given by

(∂t −∆)u = uξ (PAM)

on D = (−1, 1)3, with ξ denoting spatial white noise (constant in time), and with some

initial condition u0 ∈ Cδ(D), δ > −1/2. Fix a symmetric nonnegative ρ ∈ C∞c (R3)

integrating to 1, set ρε(x) = ε−3ρ(ε−1x), and let ξε = ρε ∗ ξ.It is then known [Hai14] that if we endow D with periodic boundary conditions and

consider the sequence of problems

(∂t −∆)v = v(ξε − Cε) , (ε-PAM)

then a suitable (diverging) choice of Cε yields a limit u independent of ρ, which we then

call “the” solution to the otherwise ill-posed problem (PAM).

First we show the existence of the Dirichlet solution of (PAM), where our result is rather

unsurprising.

Theorem 1.1. Let for ε ∈ (0, 1] uDirε be the solution to (ε-PAM) with boundary conditions

uDir

ε = 0 on R+ × ∂D.

Then for anyκ > 0,uDirε converges in probability in C([0, 1], Cδ∧(1/2)−κ(D))∩C1/2−κ

loc ((0, 1]×D) to a limit uDir independent of ρ as ε→ 0.

Page 3: Boundary renormalisation of SPDEs

Introduction 3

However, in order make sense of a notion of “Neumann solution” for (PAM), one needs

to renormalise the boundary condition as well! A similar phenomenon was previously

observed in the case of the KPZ equation in [GH19] and in the case of a stochastic

homogenisation problem in [HP21], but the boundary renormalisation remained bounded

in these examples.

Theorem 1.2. Given a constant aρ, let uε for ε ∈ (0, 1] be the solution to (ε-PAM) withboundary conditions

∂nuε = −(

aρ +|log ε|8π

)

uε on R+ × ∂D.

Then uε converges (in the same sense as in Theorem 1.1) to a limit u. Moreover, one canchoose aρ to depend on ρ in such a way that the limit does not depend on ρ.

Remark 1.3. The constant Cε is the same as in the translation invariant case, although it

will be obtained somewhat differently. It will be decomposed as Cε = ℓε( ) + ℓε( ) +4ℓε( ), with the ‘usual’ tree notation of regularity structures and ℓε denoting the BPHZ

renormalisation character, see Section 4 and the references [CH16, BHZ19]. Here ℓε( )

is of order ε−1 and ℓε( ) and ℓε( ) are of order |log ε|. This might be surprising to

readers familiar with the renormalisation of PAM: from e.g. [HP15] one would expect

instead a decomposition of the form Cε = ℓε( ) + ℓε( ). It follows from more general

identities between renormalisation constants [BGHZ19, Ger20] that these two expressions

are equivalent, although this will not play an explicit role in the present paper.

Remark 1.4. The dimension d = 3 is crucial for the boundary renormalisation to appear.

Indeed, it is easy to see that in 2 dimensions the trick in [HL15] (and described below) can

be used to solve both the Dirichlet and the Neumann problem in a quite straightforward

way. For ξ with Hölder-regularity between −4/3 and −3/2 (corresponding to “dimension”

strictly between 8/3 and 3), the setup of Section 5 allows one to apply the results of [GH19]

directly, with the translation invariant models from [CH16], resulting again in ‘standard’

statements for both the Dirichlet and the Neumann problems.

One can also justify the d = 3 threshold by simple power counting: nontrivial boundary

behaviour may only be expected if a product in the regularity structure creates a noninte-

grable singularity near the boundary, that is, of order below −1. The lowest degree product

for (PAM) with α-regular noise ξ is of order 2α + 2, which indeed shows that α = −3/2is critical.

Remark 1.5. In [BBF18] paracontrolled calculus was employed to give meaning to the

Neumann PAM. In their result however, the renormalisation in (ε-PAM) took the form of

λε (in place of Cε) for some deterministic function λε. Theorem 1.2 can be seen as giving

a precise form of this function, namely showing that one can choose λε = Cε + Cεδ∂ .

Here Cε is the same constant as in the translation-invariant case, Cε is the logarithmic

term from the theorem, and δ∂ is the ‘Dirac distribution on ∂D’ (see (2.3) for the precise

definition). Hence our solution theory is parametrised by two real constants, instead of a

space of (locally) smooth functions.

1.2 The dynamical Φ43 equation

Take the dynamical Φ43 equation

(∂t −∆)u = −u3 + ξ (Φ43)

Page 4: Boundary renormalisation of SPDEs

4 Introduction

on D = (−1, 1)3, this time ξ standing for the 1 + 3-dimensional space-time white noise,

and with initial condition u0 ∈ Cδ(D) with δ > −2/3. Consider the regularised equation

(∂t −∆)u = −u3 + 3Cεu+ ξε. (ε-Φ43)

We again start by solving the Dirichlet problem.

Theorem 1.6. Let for ε ∈ (0, 1] uDirε be the solution of (ε-Φ4

3) with boundary conditions

uDir

ε = 0 on R+ × ∂D.

Then for any κ > 0, and some random time T > 0, uDirε converges in probability in

C([0, T ], Cδ∧(−1/2)−κ(D)) to a limit uDir independent of ρ as ε→ 0.

Remark 1.7. Although uDir is a distribution, it does satisfy homogeneous Dirichlet boundary

condition in a relatively strong sense, see Remark 5.3 below.

For the Neumann problem we once again see boundary renormalisation, this time with

a twist: if one does not change the boundary conditions (i.e. bε ≡ 0 below), then the

solutions still converge, but to the Dirichlet solution uDir!

Theorem 1.8. Fix b ∈ (−∞,∞] and let (bε)ε∈(0,1] ⊂ R be such that εbε → 0 and

limε→0

|log ε|32π

− bε = b. (1.1)

Given a constant aρ, let for ε ∈ (0, 1] uε be the solution of (ε-Φ43) with boundary conditions

∂nuε = 3(aρ + bε)uε on R+ × ∂D.

Then uε converges (in the same sense as in Theorem 1.6) to a limit u. Moreover, one canchose aρ to depend on ρ in such a way that the limit u = ub only depends on b, and thisdependence is continuous. Finally, one has u∞ = uDir.

Remark 1.9. The renormalisation in the bulk is given by the same Cε as in the translation

invariant case. It is of the form Cε = ℓε( ) − 3ℓε( ), where ℓε( ) is of order ε−1 and

ℓε( ) is of order |log ε|.

1.3 The linearisation step

As usual for singular SPDEs, we start by comparing the solutions to the linear equation

with additive noise. Since this step is nontrivially affected by the boundary conditions, we

give a short outline below.

First consider the case of the PAM. Introduce, for ε ∈ [0, 1], Yε as the solution of

∆Yε = ξε on D, ∂nYε = 0 on ∂D, (1.2)

in other words,

Yε(z) =

DG(z, z′)ξε(z

′) dz′, (1.3)

where G is the Neumann Green’s function on D. Equation (1.2) of course only makes

classical sense for ε > 0, but Y0 can be simply interpreted by (1.3) in the distributional

sense.

Page 5: Boundary renormalisation of SPDEs

Introduction 5

Notice then that if u solves (ε-PAM) for some ε > 0, then v = ueYε satisfies

(∂t −∆)v = v(|∇Yε|2 − Cε) − 2∇v · ∇Yε (tPAM)

on (0, 1) ×D and v(0, x) = u0(x)eYε(x). This is essentially the same transformation that

is used in the two-dimensional case in [HL15], where it allows one to bypass the abstract

theory completely. This is not the case here, but nevertheless it will make the boundary

renormalisation easier to handle.

Remark 1.10. It is worth pointing out that the choice of boundary condition for Yε is

not related to the choice of boundary condition of (PAM), i.e. the former will always be

Neumann with vanishing data.

In contrast to Remark 1.10, for the dynamical Φ43 equation the reference linear solution

does depend of the choice of boundary conditions. Introduce, for ε ∈ [0, 1] and a ∈ [0,∞),

Ψε,a as the stationary solution of

(∂t −∆)Ψε,a = ξε on D, ∂nΨε,a = −3aΨε,a on ∂D, (1.4)

in other words,

Ψε,a(z) =

R×DG3a(z, z′)ξε(z

′) dz′, (1.5)

where G3a are the Robin heat kernels (see Section 2.3 for a detailed discussion). As before,

equation (1.4) only makes classical sense for ε > 0, but Ψ0,a can be simply interpreted

by (1.5) in the distributional sense. Recall that boundary conditions involving the normal

derivative of the solution can be incorporated on the right-hand side of the equation as a

Dirac mass on the boundary. For example, (1.5) can be equivalently written as

Ψε,a(z) =

R×DG0(z, z′)(ξε − 3aΨε,aδ∂)(z

′) dz′,

see e.g. [HP21, Rem 1.5]. It is actually not obvious that this expression also makes sense

at ε = 0. To multiply δ∂ and Ψ0,a, one would have to make sense of the restriction of the

distribution Ψ0,a to a lower dimensional subset, which is of course for a general distribution

not possible. In the case of Ψ0,a it turns out it is, see Section 3.3.

Using the above mentioned equivalence, one sees that if u solves (ε-Φ43) for ε > 0 with

boundary condition ∂nu = 3(aρ + bε)u, then v = u−Ψε,cε satisfies the equation

(∂t −∆)v = −v3 − 3v2Ψε,cε − 3v(Ψ2ε,cε −Cε − (aρ + bε + cε)δ∂)

− (Ψ3ε,cε − 3CεΨε,cε − 3(aρ + bε + cε)δ∂Ψε,cε),

(1.6)

with boundary condition ∂nv = −3cεv. This suggests that it might be a good idea to look

for cε for which the distributions Ψ2ε,cε −Cε− (aρ+ bε+ cε)δ∂ converge. This is almost the

case: in Lemma 3.2 we construct a sequence cε such that Ψ2ε,cε − ℓε( ) − (aρ + bε + cε)δ∂

converges to a ρ-independent limit.

We remark that the a = ∞ endpoint of (1.4) and (1.5) also makes perfect sense and

corresponds to the case of homogeneous Dirichlet boundary condition (see Section 2.3).

This fact will be essential in the proof of Theorem 1.8.

The rest of the article is structured as follows. In Section 2 we set up a number

of technical tools concerning function spaces and singular kernels with singularities at

Page 6: Boundary renormalisation of SPDEs

6 Preliminaries

boundaries. In Section 3 we compute explicitly the boundary renormalisation of a few

stochastic objects corresponding to the simplest ‘trees’ associated to each equation. In

Section 4 we define the regularity structures and the models associated to the equations,

which contain a few further terms that, while do not require boundary renormalisation, are

affected by the boundary conditions and therefore fall outside of the scope of the generic

construction of models [CH16]. In Section 5 we formulate each equation in the analytic

framework of [GH19], which then yields the proofs of the main results.

Acknowledgements

MG thanks the support of the Austrian Science Fund (FWF) through the Lise Meitner programme

M2250-N3 during a significant part of the project. MH gratefully acknowledges support from the

Royal Society through a research professorship. Thanks also to Etienne Pardoux for numerous

discussions on the topic of this article.

2 Preliminaries

2.1 Hölder spaces

In the computations we encounter space-time function or distribution spaces that exhibit

one or more of the following complications: they may have temporal singularities at the

initial time; they may have spatial singularities at the boundary or at the edges of the cubic

domain; they may themselves ‘live’ on the boundary of the domainD. A completely general

framework to deal with all of these seems rather cumbersome, therefore we choose to stay

in our very concrete examples.

We have two main settings: a 3-dimensional spatial and a 1+3-dimensional spatiotem-

poral one. The former is used in dealing with the stochastic objects of the PAM, while

the latter is used for the stochastic objects of the Φ43 as well as the analytic side of both

equations. We now introduce some notation that refer to different, but rather analogous

objects in the two settings.

In the spatial case (d = 3) we use the Euclidean scaling, according to which sizes

of vectors are measured and denoted by ‖ · ‖. The distance function to a given set S is

denoted by | · |S and the projection to S (wherever well-defined) is denoted by πS . The

scaled dimension therefore coincides with the usual dimension, and is also denoted by

s = 3. It will be convenient to shift the whole problem by (1, 1, 1), that is, to work

on the cube D = [0, 2]3. We denote by ∂ the boundary of D and by ∂2 the edges

of D. By symmetry, large part of the calculation can be restricted on the tetrahedron

Q = x ∈ R3 : 0 ≤ x3 ≤ x1 ≤ x2 ≤ 1 (see Figure 1), where some notations simplify:

we have |x|∂ = x3, |x|∂2 = x1, and π∂(x1, x2, x3) = (x1, x2, 0). Let us also use the

shorthand Q∂ = π∂Q = y = (y1, y2, 0) : 0 ≤ y1 ≤ y2 ≤ 1.

The analogous objects in the spatiotemporal case (d = 4) are as follows. The scaling

is parabolic, and correspondingly the scaled dimension is s = 5. Points in R4 are denoted

by (x0, x1, x2, x3) or (t, x1, x2, x3), depending on convenience. This time ∂ and ∂2 denote

the boundary and the edges of R ×D, respectively, and we set Q = x ∈ R4 : 0 ≤ x3 ≤

x1 ≤ x2 ≤ 1 as well asQ∂ = π∂Q. Similarly as before, onQ, |x|∂ = x3, |x|∂2 = x1, and

π∂(x0, x1, x2, x3) = (x0, x1, x2, 0). Below we collect a few properties of Hölder spaces

and their weighted variants.

Page 7: Boundary renormalisation of SPDEs

Preliminaries 7

x1

x2

x3

Figure 1: The tetrahedron Q (blue) and the domain D (gray).

First recall that for α < 0, the Hölder space Cα on Rd is defined as the space of

distributions u that satisfy the bounds

u(ψλy ) ≤ Cλα , (2.1)

for some C , uniformly in y over compacts, λ ∈ (0, 1], and appropriately normalised test

functions. Here and below ψλy is the rescaling (to scale λ) and recentering (around y) of ψin such a way that its integral is the same as that of ψ. More details can be found e.g. in

[Hai14, Sec 3]. Keeping in mind that the lower dimensional set ∂ admits a different scaling,

it is natural to define Cα(∂) as the set of distributions that vanish on test functions whose

support does not intersect ∂ and satisfy the bounds

u(ψλy ) ≤ Cλα−1.

The best proportionality constant C is also denoted by ‖u‖Cα(∂). Alternatively, we may

put the extra scaling factor to the test function: for any set S with a well-defined (scaled)

dimension m, we set ψλ,Sy = λs−mψλy , in which case the required bound takes the more

natural form u(ψλ,∂y ) ≤ Cλα. The definition of Cα(∂) for α ∈ [0, 1) is straightforward.

Extending the scale to α ≥ 1 is more delicate, but we do not need that generality. Note that

it follows by definition that, for distributions u supported on ∂,

‖u‖C(α∧0)−1(Rd) ≤ ‖u‖Cα(∂). (2.2)

A particular special case is the ‘Dirac distribution’ on ∂ given by

〈δ∂ , ϕ〉 :=∫

∂ϕ(x) dx. (2.3)

Page 8: Boundary renormalisation of SPDEs

8 Preliminaries

Remark 2.1. Although δ∂ is a distribution with Hölder regularity −1, it can be multiplied

with functions f well short of being C1 by setting

〈δ∂f, ϕ〉 :=∫

∂f (x)ϕ(x) dx.

This expression is perfectly meaningful if f : ∂ → R is integrable or if f : Rd → R has an

integrable trace on ∂.

For weighted Hölder spaces we take a domain S ⊂ Rd and a ‘boundary’ P ⊂ S,

with scaled codimension (as a subset of S) k. The few cases of interest to us are the

triples (S,P, k) given by (Rd, ∂, 1), (Rd, ∂2, 2), (∂, ∂2, 1), and in the spatiotemporal case

(Rd, P0, 2). Furthermore, as mentioned before, we often replace Rd by Q and ∂ by Q∂ ,

respectively. For η ≤ α ≤ 0 we define the space Cα,ηP (S) as the set of distributions on

Rd \ P that vanish on test functions whose support does not intersect S and satisfy the

bounds

u(ψλ,Sx ) ≤ Cλα|x|η−αP

uniformly in x ∈ S \ P , λ ∈ (0, 12|x|P ], and normalised test functions ψ with support in

the unit ball. For η ≤ α ∈ (0, 1) and η ≤ 0, we set Cα,ηP (S) to be the set of functions that

belong to C0,(η∧0)P (S) and satisfy the bounds

|u(x) − u(y)| ≤ C‖x− y‖α|x|η−αP

uniformly in x, y ∈ S \ P such that ‖x− y‖ ≤ 12|x|P .

The first important property of such spaces is that whenever the singularity is integrable,

that is when η > −k, they are canonically included in the space of distributions on all of Rd.

In our setting this can be formulated in the following way, which is a small modification of

e.g. [Hai14, Prop 6.9], [GH19, Prop 2.15]:

Proposition 2.2. Let (S,P, k) be as above and let −k < η ≤ α < 1. Then the spaceCα,ηP (S) canonically embeds into Cη(S) in the sense that, for every ζ ∈ Cα,ηP (S) there existsa unique ζ ∈ Cη(S) such that ζ(φ) = ζ(φ) for all test functions φ with supp φ ∩ P = ∅.

Note that in the ‘opposite’ direction one has the trivial inclusion Cη(S) ⊂ Cη,ηP (S).

When the singularity is non-integrable, there is in some cases still a natural way to obtain a

distribution on Rd.

Proposition 2.3. Let η ∈ (−2,−1). For u ∈ C0,η∂ (Q) define the distribution Ru by

(Ru)(ψ) :=

Qu(x)(ψ(x) − ψ(π∂x)) dx, ψ ∈ C∞

c (Q). (2.4)

Then the mapping u 7→ Ru is continuous from C0,η∂ (Q) to Cη(Q).

Remark 2.4. Note that while the projection πP may be not be well-defined on a set of

measure 0, the integral in (2.4) is well-defined.

Proof. Since R is linear, it suffices to show that it is bounded. Take u ∈ C0,η∂ (Q) with norm

1, and a test function of the form ψλy . We distinguish two cases, depending on whether

λ ≶ 12|y|∂ . If λ ≤ 1

2|y|∂ , then one simply has

|Ru(ψλy )| = |u(ψλy )| ≤ |y|η∂ . λη.

Page 9: Boundary renormalisation of SPDEs

Preliminaries 9

If λ ≥ 12|x|∂ , then estimating ψλy (x) − ψλy (π∂x) by |y|∂ |∇ψλy | yields the bound

|Ruε(ψλy )| .

Q∩suppψλy

|x|η∂ |x|∂λ−s−1 dy . λη.

using that 1 + η > −1 implies that |x|1+η∂ is integrable. This finishes the proof.

Remark 2.5. If u happens to be a (uniformly) smooth function, then the extension defined

by (2.4) differs from the ‘obvious’ one by δP f with some smooth function f on P . However,

u and Ru always coincide on test functions supported away from the boundary.

Multiplying weighted distributions follows the usual rules in the regularity exponent,

while the behaviour of the weight exponent can be read out from e.g. [Hai14, Prop 6.12].

Proposition 2.6. Let (S,P, k) be as above, η ≤ α ≤ 0 and η ≤ α < 1, such that α+ α > 0.Then the multiplication map is continuous from Cα,ηP (S) × Cα,ηP (S) to Cα,(α+(η∧0))∧η

P (S).

Finally, in some examples like Φ43 we have two boundaries with three different singu-

larities, one at each boundary and one at their intersection. For such a setup we take S as

before but now with two boundaries P0 and P1. We only ever encounter situations when

their codimensions are 2 and 1 respectively, and the codimension of their intersection is

3. We take a ‘weight triple’ w = (η, σ, µ), and a regularity exponent α < 1. We always

assume η, σ, µ ≤ α, as well as µ ≤ 0∧η∧σ. Then for α ≤ 0 we define the space Cα,wP0,P1(S)

as the set of distributions on Rd \ (P0 ∪ P1) that vanish on test functions whose support

does not intersect S and satisfy the bounds

u(ψλ,Sx ) ≤ Cλα|x|η−αP0|x|µ−ηP1

uniformly over λ ≤ 12|x|P0

≤ 14|x|P1

and

u(ψλ,Sx ) ≤ Cλα|x|σ−αP1|x|µ−σP0

uniformly over λ ≤ 12|x|P1

≤ 14|x|P0

.

Remark 2.7. This definition is a slight refinement of the one in [GH19, Def 4.7]. Indeed,

if we denote the (only 3-parameter) spaces therein by CwP0,P1, then as long as η > −2 and

σ > −1, Cα,wP0,P1embeds into CwP0,P1

.

For α ∈ (0, 1), we set Cα,wP0,P1(S) to be the set of functions that belong to C0,(η∧0,σ∧0,µ)

P (S)

and satisfy the bounds

|u(x) − u(y)| ≤ C‖x− y‖α|x|η−αP0|x|µ−ηP1

uniformly in x, y ∈ S \ (P0 ∪ P1) such that ‖x − y‖ ≤ 12|x|P0

≤ 14|x|P1

and the corre-

sponding symmetric bounds near P1. The following properties either follow directly from

the definition or are straightforward adaptations of some simple results of [GH19].

Proposition 2.8. Consider the above setting of S, P0, P1, w, and α.

(i) If σ > 0, then the trace function TrP1onto P1 maps Cα,wP0,P1

(S) continuously into

Cσ,η∧µP0∩P1(P1).

(ii) If η > −2, σ > −1, and µ > −3, then the space Cα,wP0,P1(S) continuously embeds into

Cη∧σ∧µ(S).

(iii) Ifα, α ∈ (0, 1) and η, σ ≤ 0, then the multiplication map is continuous from Cα,ηP0(S)×

Cα,σP1(S) to Cα,(η,σ,η+σ)

P0,P1(S).

Page 10: Boundary renormalisation of SPDEs

10 Preliminaries

2.2 Kernel bounds

Next we derive some general bounds on integrals involving singular kernels. The two

important quantities for our bounds are the scaled dimension s and the “blowup” of the

kernel that is denoted by b > 0. We are looking at a very specific blowup scenario in which

we assume

b ≤ s− 1, 2b− s = 1. (2.5)

In the two examples of the paper, we will have s = 3, b = 2 (PAM), and s = 5, b = 3(Φ4). Typically the kernels we work with are not translation invariant, which motivates

the following definition. Let G be the class of functions G : Rd × Rd → R that admit a

decomposition

G(x, y) =∑

n∈NGn(x, y),

and such that there exists a reflection T : Rd → Rd, two sets A1, A2 ⊂ R

d, and a constant

C > 0 such that:

• Gn is supported in (x, y) ∈ A1 ×A2 : ‖x− Ty‖ ≤ C2−n;

• one has the bounds |Dk1D

ℓ2G

n(x, y)| ≤ C2n(b+|k|+|ℓ|) for all x ∈ intA1, y ∈ intA2

and all multiindices k, ℓ with |k| ≤ 1, |ℓ| ≤ 1.

A trivial but important consequence of the first point is that Gn is identically 0 for 2−n .

d(x, TA1). We moreover assumeA1,A2 to be sufficiently “nice”: for our applications it will

be more than sufficient if we assume them to consist of a finite union of direct products of

intervals. If we want to emphasise the choice of parameters, we write T (G), A1(G), A2(G).

Example 2.9. Consider the 1+1-dimensional heat kernel K. The homogeneous Neumann

heat kernel on the positive half line is then given by

G(t, x, t′, x′) = K(t− t′, x− x′)1x,x′≥0 +K(t− t′, x+ x′)1x,x′≥0. (2.6)

Both terms belong to G : for the first one has T = id, A1 = A2 = R × R+, while for the

second one has T (t′, x′) = (t′,−x′), A1 = A2 = R × R+. In such a simple situation the

above formalism would be an overkill, and some of the calculations below were actually

performed in [GH19]. In higher dimensions however, there are several different reflections

that need to be handled, hence the more generic setup of G .

For G ∈ G we define

Gε(x, y) =

G(x, z)ρε(y − z) dz. (2.7)

We need three basic bounds for kernels of this type. We remark the elementary bounds, for

δ1, δ2 > 0, α1 < 0 < α2,

n∈N: 2−n.δ1

2nα1 . δ−α1

1 ,∑

n∈N: 2−n&δ2

2nα2 . δ−α2

2 ,

that are repeatedly used in the proofs.

Page 11: Boundary renormalisation of SPDEs

Preliminaries 11

Lemma 2.10. Let G, G ∈ G with A1 = A1 and such that (2.5) holds. Take γ ∈ (0, 1).Then for all x, x′ ∈ A1 with |x− x′| . d(x, T A2) one has

GεGε(x, y) dy . |ε ∨ d(x, T A2)|−1; (2.8)

GεGε(x, y) −GεGε(x′, y) dy . ‖x− x′‖γ |ε ∨ d(x, T A2)|−1−γ . (2.9)

Proof. Let us use the shorthand [x] = d(x, T A2). We only prove the second bound since

the first one is easier. First note that one has the decomposition

Gε(x, y) =∑

n∈NGnε (x, y),

where supp Gnε (x, ·) ⊂ y : ‖x− T y‖ . ε+ 2−n, d(y, A2) ≤ ε and one has bounds

|Gnε (x, y)| . (ε ∨ 2−n)−s2n(b−s)

|Gnε (x, y) − Gnε (x′, y)| . |x− x′|γ(ε ∨ 2−n)−s2n(b−s+γ)

for all x, x′ ∈ A1 and y ∈ Rd, and similarly for G. In particular, the volume of (suppGnε )∩

(supp Gmε ) is bounded by (ε ∨ (2−n ∧ 2−m))s. Therefore, the left-hand side of (2.9) is

bounded by

|x− x′|γ∑

n,m∈N(ε ∨ (2−n ∧ 2−m))s(ε ∨ 2−n)−s2n(b−s)

× (ε ∨ 2−m)−s2m(b−s)(2nγ ∨ 2mγ)

= |x− x′|γ∑

n,m∈N2(n+m)(b−s)(ε ∨ 2−n ∨ 2−m)−s(2nγ ∨ 2mγ)

(2.10)

At this stage the roles of n and m are symmetric, so we can bound the above sum by∑

2−n&2−m

2(n+m)(b−s)(ε ∨ 2−n)−s2mγ .∑

n∈N2n(2b−2s+γ)(ε ∨ 2−n)−s

. ε−s∑

2−n.ε

2n(2b−2s+γ) +∑

2−n&ε

2n(2b−s+γ) . ε−1−γ ,(2.11)

where we used that b− s+γ < 0 and 2b− s = 1. This yields the required bound if [x] . ε.For [x] & εwe make use of the property that Gmε is identically 0 for 2−m . [x]. Therefore,

instead of (2.10), we now get the bound

|x− y|γ∑

2−m&[x]n∈N

2(n+m)(b−s)(2−n ∨ 2−m)−s(2nγ ∨ 2mγ).

For the part of the sum where 2−n & 2−m, we get a bound∑

2−n&2−m&[x]

2nb2m(b−s+γ) .∑

2−m&[x]

2m(2b−s+γ) . [x]−1−γ ,

as required. Concerning the 2−m & 2−n regime, we can write∑

2−m&(2−n∨[x])

2mb2n(b−s+γ) .∑

2−m&[x]

2m(2b−s+γ) . [x]−1−γ ,

which finishes the proof.

Page 12: Boundary renormalisation of SPDEs

12 Preliminaries

Lemma 2.11. Assume the setting of Lemma 2.10. Then for all x ∈ A1 with d(x, T A2) & ε,one has

(GεGε −GG)(x, y) dy∣

∣. εγ(d(x, T A2))−1−γ . (2.12)

Proof. We writeGεGε−GG = (Gε−G)G+Gε(Gε−G) and bound the two corresponding

integrals separately (unfortunately the two cases are not exactly symmetric). We first treat

(Gε −G)G. Writing Gn∆ = Gnε −Gn, the quantity to bound is

m,n∈N

(Gn∆Gm(x, y)) dy.

Using again the shorthand [x] = d(x, T A2), we see that the sum overm can be restricted to

the range 2−m & [x], since Gm vanishes otherwise. The easiest case is 2−n . ε, one can

simply use the bounds |suppGn∆| . 2−ns, sup |Gn∆| . 2nb, sup |Gm| . 2mb. This yields

2−m&[x]

2−n.ε

(Gn∆Gm(x, y)) dy

∣.

2−m&[x]

2−n.ε

2n(b−s)2mb . εs−b[x]−b.

Recalling that [x] & ε, s ≥ b + 1, and 2b − s = 1, one sees that this is indeed bounded

by ε[x]−2, as required. For 2−n & ε we split the integral into two regions, depending on

the distance of y to ∂A2. Define ∂ε = y : d(y, ∂A2) . ε. If y /∈ ∂ε, then one can use

the differentiability of G in the second variable to get the bound |Gn∆(x, y)| . εγ2n(b+γ).

Therefore,

2−m&[x]

2−n&ε

(∂ε)c(Gn∆G

m(x, y)) dy∣

∣.

2−m&[x]

2−n&ε

(2−ns ∧ 2−ms)εγ2n(b+γ)2mb. (2.13)

For the regime 2−n & 2−m one has

2−n&2−m&[x]

2−msεγ2n(b+γ)2mb . εγ∑

2−m&[x]

2m(2b−s+γ) . εγ[x]−1−γ .

For 2−m & 2−n & [x] one gets

2−m&2−n&[x]

2−nsεγ2n(b+γ)2mb . εγ∑

2−m&|x|2m(2b−s+γ) . εγ[x]s−2b−γ = εγ[x]−1−γ .

Finally, in the case 2−n . [x] the sum becomes

2−m&[x]&2−n&ε

2−nsεγ2n(b+γ)2mb . εγ[x]−b∑

[x]&2−n

2n(b+γ−s) . εγ[x]s−2b−γ .

All of these bounds are of the required order. It remains to treat the ∂ε portion of the

integral. Note that in this case the size of the region of integration is at most of order

ε(2−n(s−1) ∧ 2−m(s−1)). Combining this with the trivial supremum bounds one gets

2−m&[x]

2−n&ε

∂ε(Gn∆G

m(x, y)) dy∣

∣.

2−m&[x]

2−n&ε

ε(2−n(s−1) ∧ 2−m(s−1))2nb2mb. (2.14)

Page 13: Boundary renormalisation of SPDEs

Preliminaries 13

Again, first bound the sum over 2−n & 2−m:

2−n&2−m&[x]

ε2−m(s−1)2nb2mb . ε∑

2−m&[x]

2m(2b−s+1) . ε[x]−2,

as required. Next, in the case [x] . 2−n . 2−m one has

2−m&2−n&[x]

ε2−n(s−1)2nb2mb . ε∑

2−n&[x]

2n(2b−s−1) . ε[x]−2,

as required. Finally, for 2−n . [x] the sum becomes

2−m&[x]&2−n&ε

ε2−n(s−1)2nb2mb . ε[x]−b∑

[x]&2−n&ε

2n(b−s+1)

. εγ[x]−b∑

[x]&2−n&ε

2n(b−s+γ) . εγ[x]−1−γ ,

using that b− s+ γ < 0. Combining all the cases finishes the term (Gε −G)G.

It now remains to do a similar calculation for Gε(Gε − G). Writing Gm∆ = Gmε −Gm,

the quantity to bound is∑

m,n∈N

(Gnε Gm∆(x, y)) dy.

As before, the sum over m can be restricted to the regime 2−m & [x], since both Gm and

Gmε vanish otherwise. For the former this is obvious and for the latter this follows from the

assumption ε . [x]. To bound the sum over 2−n . ε, one can use ‖Gnε ‖L1 = ‖Gn‖L1 .

2n(b−s), with the trivial bound sup |Gm∆ | . 2mb. This yields the same bound as before,

namely

2−m&[x]

2−n.ε

(Gnε Gm∆(x, y)) dy

∣.

2−m&[x]

2−n.ε

2n(b−s)2mb . εs−b[x]−b.

For 2−n & ε we split the integral to ∂ε and (∂ε)c as before, and use that if y /∈ ∂ε, then

one due to the differentiability of G in the second variable one has the bound |Gm∆ (x, y)| .εγ2m(b+γ). Therefore,

2−m&[x]

2−n&ε

(∂ε)c(Gnε G

m∆(x, y)) dy

∣.

2−m&[x]

2−n&ε

(2−ns ∧ 2−ms)2nbεγ2m(b+γ).

We leave it as an exercise to the reader to treat this sum similarly to the one in (2.13). Finally,

concerning the integral over ∂ε we get

2−m&[x]

2−n&ε

∂ε(Gnε G

m∆(x, y)) dy

∣.

2−m&[x]

2−n&ε

ε(2−n(s−1) ∧ 2−m(s−1))2nb2mb.

The right-hand side is now precisely the same as in (2.14), and therefore using the already

established bound the proof is finished.

Page 14: Boundary renormalisation of SPDEs

14 Preliminaries

Lemma 2.12. Let G ∈ G , let A ⊂ Rd be an open convex set, and define G(x, y) =

G(x, y)1y∈A. Take γ ∈ (0, 1), γ′ ∈ [−1, 1]. Then for all x, x′ with ‖x − x′‖ . d(x, ∂A)

one has∣

(

G2ε(x, y) − G2

ε(x′, y))

−(

G2ε(x, y)1y∈A −G2

ε(x′, y)1y∈A)

dy∣

. ‖x− y‖γεγ′(d(x, ∂A))−1−γ−γ′ .

(2.15)

(Recall the notation introduced in (2.7).)

Proof. Let us again use the shorthand [x] = d(x, ∂A). Clearly it suffices to consider the

extremal cases γ′ ∈ −1, 1. For γ′ = −1 we can bound the integrals of the two terms

in the big brackets as in (2.10)-(2.11). In fact this gives the required bound not only for

γ′ = −1, but also for γ′ = 1 in case [x] . ε.In the case γ′ = 1, [x] & ε, the integrand in (2.15) vanishes identically on [y] & ε.

On the remaining region we again bound the integrals of the two terms in the big brackets

separately. They are essentially identical calculations, we only detail the first one. One can

write∫

[y].εG2ε(x, y) − Gε(x

′, y) dy

. ‖x− x′‖γ∑

n,m∈N

∣(supp Gnε ) ∩ (supp Gmε ) ∩ [y] . ε

∣2nb2mb(2nγ ∨ 2mγ).

Since [x] & ε, only terms with 2−n, 2−m & [x] contribute to the sum. In this case one has∣

∣(supp Gnε ) ∩ (supp Gmε ) ∩ [y] . ε

∣. ε(2−n ∧ 2−m)s−1.

Since the above sum is symmetric under n↔ m, we can bound it by

ε∑

2−m&2−n&[x]

2n(b+γ+1−s)2mb . ε∑

2−n&[x]

2n(2b+γ+1−s) . ε[x]−2−γ ,

as required.

2.3 Convergence of the Robin kernels

The purpose of this section is to make the folklore fact

“the −∞ Robin boundary condition is the 0 Dirichlet boundary condition”

precise in a form that suits the setup of [GH19]. The dimension plays no role here, but we

stick to the 1 + 3-dimensional setting that we will use later. As a warm-up example, let us

recall the construction of Robin heat kernels on the upper half spaceR4u = (x0, x1, x2, x3) :

x3 > 0. For x ∈ R4, r ∈ R denote xr = (x0, x1, x2,−x3−r). Then for any a ∈ (−∞,∞)

the function

Ga(x, y) = K(x− y) +K(x− y0) −∫ ∞

0

2ae−arK(x− yr) dr (2.16)

is easily seen to satisfy, for each fixed y ∈ R4u,

(∂x0 −∆x)Ga(x, y) = δx=y on R4u;

Page 15: Boundary renormalisation of SPDEs

Preliminaries 15

∂nGa(x, y) = −∂x3 Ga(x, y) = −aGa(x, y) on ∂R4u.

Therefore Ga – more precisely, its product with the indicator of (R4u)2 – is indeed the Robin

heat kernel. Taking a = ∞ in (2.16) the measure ae−ar dr becomes the Dirac mass at 0,

and we recover the Dirichlet heat kernel. Loosely speaking, to build the Robin kernels on

the cube D, one needs to repeat the procedure of reflecting and averaging in (2.16) for each

face of the cube ad infinitum.

Remark 2.13. Note that for a ∈ [0,∞], the function Ra(x, y) = 1x,y∈R4u

∫∞0

2ae−arK(x−yr) dr is twice a convex combination of the kernels 1x,y∈R4

uK(x − yr) that each fit in the

framework of Section 2.2 withA1 = A2 = R4u and T (y) = yr (hence also with T (y) = y0).

Therefore Ra satisfy the bounds therein uniformly over a ∈ [0,∞].

To formulate the result, recall the following concept of kernel remainders from [GH19].

For the present section it is more convenient to work on the cube D = (−1, 1)3 and recall

that we denote by ∂ the boundary of R×D.

Definition 2.14. Denote by Zβ,∂ the set of functions Z : (R4 \∂)2 → R that can be written

in the form Z(z, z′) =∑

n≥0 Zn(z, z′) where, for each n, Zn satisfies the following

• Zn is supported on (z, z′) = ((t, x), (t′, x′)) : |x|∂ + |x′|∂ + |t− t′|1/2 ≤ 3(2−n),

where C is a fixed constant depending only on the domain D.

• For any multiindices k and ℓ with |k|, |ℓ| ≤ 2,

|Dk1D

ℓ2Zn(z, z′)| . 2n(s+|k+ℓ|−β), (2.17)

where the proportionality constant may depend on k and ℓ, but not on n, z, z′.

Clearly Zβ,∂ is a vector space, on which the best proportionality constant in (2.17)

defines a norm ‖·‖Zβ,∂. Let us decompose the 1+3-dimensional heat kernel as K = K+R

in such a way that K = K on the ball of radius 1/2 around the origin and vanishes outside

the ball of radius 1. Furthermore, K can be chosen to satisfy [Hai14, Ass. 5.1] and R is

globally smooth with any derivatives having faster than polynomial spatial decay.

Whenever a ∈ (−∞,∞)\0, the boundary conditions (∂na +1)f = 0 and (∂n+a)f =0 are equivalent. For a = 0 only the latter makes sense (and gives the homogeneous

Neumann boundary conditions), while for a = ∞ only the former does (and gives the ho-

mogeneous Dirichlet). In the lemma below we use the former form, with the understanding

of the obvious modification for a = 0.

Lemma 2.15. There exists a family of remainders (Z (a))a∈(−∞,∞] such that:

(i) (∂t −∆)(K + Z (a))(t, x, t′, y) = δt=t′,x=y on ([0, 1] ×D)2;

(ii) (∂na + 1)(K + Z (a))(t, x, t′, y) = 0 on [0, 1] × ∂D × [0, 1] ×D;

(iii) For all β < 2, (Z (a))a∈(−∞,∞] is continuous with respect to the natural topology of(−∞,∞], as a function with values in Zβ,∂ .

We then denote the Robin heat kernels on D by

Ga = K + Z (a). (2.18)

Page 16: Boundary renormalisation of SPDEs

16 Preliminaries

Proof. Let us preface that since the continuity property is only easier in (−∞,∞), we will

only deal with it at the endpoint case a = ∞. Denote by S the group of transformations of

R3 generated by the reflections g±i on the hyperplanes R

i−1 × ±1 × R3−i, i = 1, 2, 3,

and the different pieces of the boundary by D±i := (−1, 1)i−1 ×±1 × (−1, 1)3−i. Also

let e±i denote the outward normal vector on the boundary piece D±i.Take a ∈ [1,∞] and a function F : (R4)2 → R with all derivatives having faster than

polynomial spatial decay, which furthermore has a ‘sign’ b ∈ ±13 with the property

∂xiF (t, x, t′, y) = bi∂yiF (t, x, t′, y). Let us denote by T a±iF the function

(T a±iF )(t, x, t′, y) =

∫ ∞

0

bi(−1 + 2e−as)∂e±iF (t, x, t′, g±i(y) + se±i) ds.

Here and below the partial derivative ∂e±i is understood to act onto the x coordinate. Note

that T a±iF also has the above mentioned properties, with its ‘sign’ switched in the i-thcoordinate. The construction is such that one has

(∂±eia + 1)(F + T a±iF ) = 0 on R×D±i × R× R

3. (2.19)

Notice also that for i 6= j, one has T±iT±j = T±jT±i, and so for any g ∈ S for any

minimal (with respect to the length) representation g = h1h2 · · · hn, where hk = g±i, the

corresponding mapping T ag = T ah1 · · ·T ahn is well-defined. We also write T aid = id.

Let A be the set of elements of S whose minimal representation contains at most one

of g−i and g+i for all i. Consider

G1,a :=∑

g∈AT ag K.

Note that for any fixed ±i, all g ∈ A are of one of three types: not containing either g−ior g+i for all i, of the form g±ig with g of the previous type, or containing g∓i. Since for

those of the last type, (T ag K)(t, ·, t′, y) vanishes near D±i whenever y ∈ D, one can ‘pair

up’ elements of the first two types, and conclude by (2.19) that

(∂±eia + 1)G1,a = 0 on R×D±i ×R×D.

Fix β < 2, for convenience and without loss of generality we also assume β ≥ 1. Let us

use the notation f ∼ f for functions f , f on (R4)2 whenever f = f on ([0, 1] ×D)2. Note

for instance, one has T ag K ∼ 0 for all g ∈ S \ A. We then claim that for all g ∈ A \ id,

there exists a Zag ∈ Zβ,∂ such that T ag K ∼ Zag , and ‖Zag − Z∞g ‖Zβ,∂

→ 0 as a → ∞. For

convenience let us illustrate the argument for g = g+i: first note that ifx, y ∈ D are such that

(t, x, t′, g+i(y)+se+i) ∈ supp Kn for some t, t′, s ≥ 0, then d(x, ∂D)∨d(y, ∂D) ≤ 2−n−1.

Take now smooth functions ϕn on (R3)2, which are 1 on (x, y) : d(x, ∂D) ∨ d(y, ∂D) ≤2−n−1, supported on (x, y) : d(x, ∂D) ∨ d(y, ∂D) ≤ 2−n, and for all multiindices kand ℓ, Dk

1Dℓ2ϕ is bounded by 2n(|k+ℓ|), up to a constant uniform in n. One then has, on

([0, 1] ×D)2,

T ag+iK =

n≥0

∫ ∞

0

−(−1 + 2e−as)∂e+iKn(t, x, t′, g+i(y) + se+i) ds

∼∑

n≥0

ϕn(x, y)

∫ ∞

0

−(−1 + 2e−as)∂e+iKn(t, x, t′, g+i(y) + se+i) ds

Page 17: Boundary renormalisation of SPDEs

Explicit boundary corrections 17

=: Zag+i= Z∞

g+i+ (Zag+i

− Z∞g+i

).

Noticing that the values of s with non-zero contribution to the integral above are of size at

most O(2−n), one has

2−n(s+|k+ℓ|−β)|Dk1D

ℓ2(Zag+i

− Z∞g+i

)n| . 2n(β−1)

∫ C2−n

0

e−as ds

≤ 2n(β−1)(2−n ∧ a−1) ≤ aβ−2 , (2.20)

which indeed converges to 0 uniformly in n, as a→ ∞, yielding our claim.

Consider next

G2,a := ϕ0

g∈ST agR.

Thanks to the spatial decay properties of R, this sum converges as a smooth function on

([0, 1] × D)2, uniformly in a ∈ [0,∞]. In particular, this is enough to infer from (2.19)

that it also satisfies the boundary condition, and also that since for each g, ‖ϕ−1Tag R −

ϕ−1T∞g R‖Zβ,∂

→ 0, one has ‖G2,a −G2,∞‖Zβ,∂→ 0.

Since clearly (∂t −∆)(T ag (K +R)) = 0 on ([0, 1] ×D)2 for g 6= id, setting

Z (a) := G2,a +∑

g∈A\idZag ,

completes the proof of the lemma.

Remark 2.16. As can be seen from the explicit expression (2.16), the limit a → −∞ does

not exist, contrarily to what (ii) may suggest.

Remark 2.17. In the special cases a = 0,∞ the above construction coincides with the one

in [GH19, Ex. 4.15] for the homogeneous Neumann and Dirichlet heat kernels, respectively.

3 Explicit boundary corrections

In this section we perform the boundary renormalisation of expectations of some concrete

stochastic objects. Recall that in the simplest situation of translation invariant renormalisa-

tion without subdivergences one considers a sequence of random distributions Xε, where

EXε = Cε diverges but does not depend on the space-time variable. In the case with

boundaries the prototypical situation will be

EXε = Cε + cεδ∂ + R1ε + R2

ε + R3ε, (3.1)

with divergent Cε and cε and three different types of remainder distributions Riε that each

converge to a finite limit.

3.1 PAM

We denote Ψε = ∇Yε, where Yε is defined in (1.2). Let K(x) = 14π|x| be the Green’s

function of the 3-dimensional Poisson equation and fix a compactly supported function Ksuch that K − K is smooth and vanishes in a neighbourhood of the origin. Set

ℓε( ) = E|∇K ∗ ξε|2. (3.2)

Page 18: Boundary renormalisation of SPDEs

18 Explicit boundary corrections

As mentioned in Remark 1.3, the quantity ℓε( ) arises from the BPHZ renormalisation of

the regularity structure associated to (PAM). We will give more details in Section 4 below,

but the reader may freely take (3.2) as a definition for now.

We will show a decomposition of the type (3.1) on the tetrahedron Q.

Lemma 3.1. For all ε > 0, on Q one has the decomposition

E|Ψε|2 = ℓε( ) + (aρ +|log ε|8π )δ∂ + δ∂R

1ε + RR2

ε +R3ε,

where aρ is a constant, and the remainders satisfy:

(i) R1ε → R1

0 in C1−κ,−κ∂2

(Q∂);

(ii) R2ε → R2

0 in C1−κ,−1−κ∂ (Q) and Dx1R

2ε = Dx2R

2ε = 0;

(iii) R3ε → R3

0 in C1−κ,−1−κ∂2

(Q);

where κ > 0 is arbitrarily small and the limits do not depend on ρ.

Proof. Introduce the shorthand, for functions f on (R3)2, ε ∈ [0, 1], and S ⊂ R3,

(DεSf )(x, x′) := (ρε ∗ (∇xf (x, ·)1S(·)))(x′)

with the convention that for ε = 0 we replace the convolution with ρε by the identity. With

this notation we have

ℓε( ) =

|DεR3K(x, x′)|2 dx′,

which of course does not actually depend on x since K depends only on the difference of

its arguments. One can then write, for ε > 0,

EΨ2ε(x) − ℓε( ) =

(|DεDG(x, x′)|2 − |Dε

R3K(x, x′)|2) dx′.

As a first step, we truncate the infinite sum in G and remove the truncation of K . Let

B = ±13, and for b ∈ B and x ∈ R3, let xb be the vector obtained by switching the signs

of the coordinates of x according to b. Denote furthermore Kb(x, x) := K(x, xb). We then

write

EΨ2ε(x) − ℓε( ) −R3,1

ε (x) =

(∣

∣DεD

(

b∈BKb

)

(x, x′)∣

2

− |DεR3K(x, x′)|2

)

dx′, (3.3)

interpreting this as a definition of R3,1ε . It follows from the reflection principle that, on Q,

R3,1ε converges as a smooth function to a ρ-independent limit. Next we claim that with

R3u = (x1, x2, x3) : x3 > 0,

R3,2ε (x) =

∣DεD

(

b∈BKb

)

(x, x′)∣

2

− |DεR3u(K (1,1,1) +K (1,1,−1))(x, x′)|2 dx′

converges in C1−κ,−1−κ∂2

(Q) to a ρ-independent limit. The function R3,2ε can be written as

a finite linear combination of terms of the type

Rε(x) :=

(DεSK

b ·DεSK b)(x, x′) dx′ ,

Page 19: Boundary renormalisation of SPDEs

Explicit boundary corrections 19

where either b ∈ B \ (1, 1, 1), (1, 1,−1) and S = D, or b ∈ (1, 1, 1), (1, 1,−1) and

S = R3u \D. The choice of b and S will not play a role. Notice that we are in the setting

of Section 2.1: ∇xKb(x, y)1y∈S ∈ G , with s = 3, b = 2, T (z) = zb, Ax = R

3, Ay = S.

Moreover, for x ∈ Q, one has d(x, S b) ≥ x1 = |x|∂2 for each choice of b and S as above,

and therefore Lemma 2.10 provides a bound for Rε in C1−κ,−1

∂2(Q), uniformly in ε. Since

on Q \ ∂2, Rε converges locally in C1−κ, this proves the convergence in C1−κ,−1−κ∂2

(Q) for

each Rε, and consequently for R3,2ε as well. The function R3

ε = R3,1ε + R3,2

ε therefore

satisfies (iii) and we have so far proved the following decomposition on Q:

EΨ2ε(x)− ℓε( )−R3

ε(x) =

|DεR3u

(

K (1,1,1) +K (1,1,−1))

(x, x′)|2 − |DεR3K(x, x′)|2 dx′.

Define R2ε as the right-hand side of the above equality. It is clear that R2

ε does not depend

on x1 and x2, so it remains to check its convergence in C1−κ,−1−κ∂ (Q). By Lemma 2.12,

both of the functions∫

|DεR3uK (1,1,1)(x, x′)|2 − |Dε

R3K(x, x′)|21x′∈R3udx′,

|DεR3uK (1,1,−1)(x, x′)|2 − |Dε

R3K(x, x′)|21x′∈(R3\R3u) dx

′,

converge to 0 in C1−κ,−1−κ∂ (Q). Concerning the cross term in R2

ε , its convergence to

R20(x) = 2

1x′∈R3u(∇1K · ∇1K

(1,1,−1))(x, x′) dx′

follows as above: Lemma 2.10 yields a uniform bound in C1−κ,−1∂ (Q), and away from ∂ the

convergence in C1−κ is quite clear. Therefore, R2ε satisfies (ii). It remains to show that the

difference R2ε − RR2

ε is of the claimed form.

As noted in Remark 2.5, this difference is of the form δ∂mε, and on Qδ one can express

the function mε by

mε(y) =

∫ y1

0

R2ε(y1, y2, s) ds =

∫ y1

0

R2ε(0, 0, s) ds.

Let us use the shorthand R2ε(0, 0, s) = Iε(s), for which we have the bound ε−1 from (2.10).

One can rewrite the above integral as

mε(y) =

∫ ε

0

Iε(s) ds+

∫ ∞

εIε(s) − I0(s) ds −

∫ ∞

y1

Iε(s) − I0(s) ds+

∫ y1

εI0(s) ds.

By Lemma 2.11 and 2.12, one has the bound |Iε(s) − I0(s)| . ε1−κ/s2−κ. Therefore, the

second term above is finite and by scaling invariance, is independent of ε, so it is just a

(ρ-dependent) constant. The first term is also independent of ε, also by scaling invariance.

Therefore,

mε(y) = aρ +

∫ ∞

y1

Iε(s) − I0(s) ds+

∫ y1

εI0(s) ds.

Denote the second term by R1,1ε . Invoking Lemma 2.11 again, we have |R1,1

ε (y)| ≤ε1−κ/y1−κ1 , and by (2.8) we have ∇yR

1,1ε (y) ≤ 1/y1. This is enough to conclude R1,1

ε → 0

Page 20: Boundary renormalisation of SPDEs

20 Explicit boundary corrections

in C1−κ,−κ∂2

(Q∂). Moving on to the third term on the right-hand side we write, with

s = (0, 0, s),

I0(s) =

R3

(∇1K · ∇1K(1,1,−1))(s, x′) dx′

=

R3

((−∂x′1,−∂x′

2,−∂x′

3)K · (−∂x′

1,−∂x′

2, ∂x′

3)K (1,1,−1))(s, x′) dx′

= −∫

R3

((∆K)K (1,1,−1))(s, x′) dx′ − 2

R3

(∂x′3K∂x′

3K (1,1,−1))(s, x′) dx′

=: I10 (s) +1

16π2I20 (s).

Since −∆K = δ0, one easily gets

I10 (s) =

R3

δx′=sK(1,1,−1)(s, x′) dx′ =

1

8πs.

Next we rewrite I20 (s) by change of variables: first by setting xi = x′i/s and then x3 = 1/x3,x1 = x1/x3, x2 = x2/x3, one gets

I20 (s) =

R3u

(s− x′3)(s + x′3)

|(x′1, x′2, x′3 − s)|3|(x′1, x′2, x′3 + s)|3 dx′

=1

s

R3u

1− x23|(x1, x2, x3 − 1)|3|(x1, x2, x3 + 1)|3 dx

=1

s

R3u

x23 − 1

|(x1, x2, x3 − 1)|3|(x1, x2, x3 + 1)|3 dx = −I20 (s),

and so I20 (s) = 0. We can conclude that

mε(y) = aρ +R1,1ε (y) + log y1

8π − log ε8π ,

and, setting R1ε(y) = R1,1

ε (y) + log y18π , this completes the proof.

3.2 Φ43 - the quadratic term

Let K be the heat kernel on R × R3 and fix a compactly supported function K such that

K − K is smooth and vanishes in a neighborhood of the origin. We then define

ℓε( ) = E(K ∗ ξε)2.

As before, we will show a decomposition of the type (3.1) on Q.

Lemma 3.2. Let (bε)ε∈(0,1] ⊂ R be a sequence such that εbε → 0 and

limε→0

( |log ε|32π

− bε

)

= b ∈ [0,∞].

Then there exist a sequence (cε)ε∈(0,1] such that cε → b and such that on Q one has thedecomposition

E|Ψε,cε|2 = ℓε( ) + (aρ + bε + cε)δ∂ + δ∂R1ε + RR2

ε +R3ε,

where aρ is a constant and the remainders Riε satisfy the properties in Lemma 3.1 (i)-(iii).

Page 21: Boundary renormalisation of SPDEs

Explicit boundary corrections 21

Remark 3.3. While for b < ∞ one may take cε ≡ b, for b = ∞ the sequence cε is notobtained in the trivial way cε = |log ε|

32π − bε. For example, when bε ≡ 0, the difference

cε− |log ε|32π should actually be chosen to diverge at order log |log ε|. This is left as an exercise

to the interested reader.

Proof. Let us first take an arbitrary sequence cε → b such that εcε → 0. The first part of

the argument is then virtually identical to that in the proof of Lemma 3.1. By following the

same steps, we can conclude that on Q one has

EΨ2ε,cε − ℓε( ) −R3

ε − RR2ε = δ∂mε

with R2ε and R3

ε satisfying (ii) and (iii), respectively. It is also clear that R20 and R3

0 do not

depend on cε but only on b (since they can be expressed from heat kernels for the −3b-Robin

boundary condition). The function mε is given on Q∂ by

mε(y) =

∫ y1

0

|(ρε ∗ G3cε(s, ·))(x)|2 − |(ρε ∗ K(s, ·))(x)|2 dx ds, (3.4)

where s = (0, 0, 0, s) and G3cε is the Robin heat kernel on the upper half space R4u =

(x0, x1, x2, x3) : x3 > 0. Recall from (2.16) that it is given, with the notation zr =(z0, z1, z2,−z3 − r), by

Ga(z, z) = 1z3≥0K(z − z) + 1z3≥0K(z − z0) −Ra(z, z)

= 1z3≥0K(z − z) + 1z3≥0K(z − z0) − 1z3≥0

∫ ∞

0

2ae−arK(z − zr) dr.

Now we would like to proceed similarly to Lemma 3.1 by simplifying the integral in

(3.4). However, since the kernels G3cε themselves depend on ε, some of the scaling

arguments break down. Therefore let us separate the ε-dependent part from the kernel.

Let Kε(z, z) = (ρε ∗ (K(z − ·)))(z), Kε+(z, z) = (ρε ∗ (1·∈R4

uK(z − ·)))(z), Kε

−(z, z) =(ρε ∗ (1·∈R4

uK(z − (·)0)))(z), and Rε

a(z, z) = (ρε ∗ (Ra(z, ·)))(z). and define

J εa (s) =

R4u

−2Kε+(s, z)Rε

3a(s, z) − 2Kε−(s, z)Rε

3a(s, z) + |Rε3a(s, z)|2 dz. (3.5)

We then have

mε(y) =

∫ y1

0

|(Kε+ +Kε

−)(s, z)|2 − |Kε(s, z)|2 dz ds +∫ y1

0

J εcε(s) ds.

Now the first integral can be treated by the scaling arguments as in Lemma 3.1. That is,

there exists a constant aρ and a sequence of functions R1,1ε converging to 0 in C1−κ,−κ

∂2(Q∂)

such that

mε(y) = aρ +R1,1ε (y) +

∫ y1

εI0(s) ds +

∫ y1

0

J εcε(s) ds,

where the function I0 is given by

I0(s) =

R4u

2K(s − z)K(s − z0) dz.

Page 22: Boundary renormalisation of SPDEs

22 Explicit boundary corrections

By scaling invariance once again, we have I0(s) = 1sI0(1). The value of I0(1) can be found

by an explicit computation, which we perform in a more general setting below, see (3.15),

which yields the value

I0(1) =1

32π. (3.6)

Therefore,

mε(y) = aρ +|log ε|32π +R1,1

ε (y) + log y132π +

∫ y1

0

J εcε(s) ds. (3.7)

Clearly, R1,2ε (y) := log y1

32π satisfies (i). Moving on the last term, we rewrite it as

∫ ε

0

J εcε(s) ds +

∫ y1

ε(J ε

cε(s) − J 0cε(s)) ds−

∫ 1

y1

J 0cε(s) ds +

∫ 1

εJ 0cε(s) ds. (3.8)

By the scaling relation Rεa(z, z) = Rλε

λ−1a(λz, λz)λ3, we have J εa (s) = J λε

λ−1a(λs)λ and

therefore∫ ε

0

J εcε(s) ds =

∫ 1

0

J 1εcε(s) ds . (3.9)

DenoteKεr(z, z) = (ρε∗(1·∈R4

uK(z−(·)r)))(z), which is just the generalisation of Kε

− = Kε0.

From (2.8) one has the bounds∣

K1+(s, z)K1

r (s, z) dz∣

∣,∣

K1r′(s, z)K1

r (s, z) dz∣

∣. 1 ∧ r−1

uniformly in s, r, r′ ≥ 0. Since εcε → 0, this implies∫ 1

0J 1εcε

(s) ds → 0. Denote the

second integral in (3.8) byR1,3ε (y). From Lemma 2.11 one has the bound |J ε

a (s)−J 0a (s)| .

ε1−κ/s2−κ, uniformly in a (see Remark 2.13). Therefore, if similarly to (3.9) we write

R1,3ε (y) =

∫ y1

ε(J ε

cε(s) −J 0cε(s)) ds =

∫ ε−1y1

1

(J 1εcε(s) − J 0

εcε(s)) ds,

then by the dominated convergence theorem, the right-hand side goes to 0. Moreover, one

has the bound Dy1R1,3(y) . 1/y1 from (2.8). Therefore R1,3

ε → 0 in C1−κ,−κ∂2

(Q∂). Next,

denote the third integral in (3.8) by R1,4ε (y). Its convergence away from y1 = 0 is clear,

and therefore a uniform in ε bound in C1,0∂2

(Q∂) suffices to conclude the property (i). This

is quite immediate: R1,4ε vanishes on the hyperplane y1 = 1, and one has the bound

Dy1R1,3ε (y) . 1/y1 from (2.8) as before.

So R1ε = R1,1

ε +R1,2ε +R1,3

ε +R1,4ε satisfies (i), and one can rewrite (3.7) as

mε(y) = aρ +|log ε|32π +R1

ε(y) +

∫ 1

εJ 0cε(s) ds

= aρ +|log ε|32π +R1

ε(y) +

∫ 1

ε

1

sJ 0scε(1) ds.

We are therefore finished as soon as we show that there exist solutions cε, dε to

bε + cε + dε =|log ε|32π +

∫ 1

ε

1

sJ 0scε(1) ds

that furthermore satisfy cε → b, εcε → 0, and dε → d0 for some finite and ρ-independent

d0. We claim the following properties of the function J 0a (1), whose proof we postpone so

that the present proof can be concluded.

Page 23: Boundary renormalisation of SPDEs

Explicit boundary corrections 23

Proposition 3.4. The function a 7→ J 0a (1) is continuous on (0,∞), lima→∞ J 0

a (1) =−2I0(1), and the bound

|J 0a (1)| . a|log a| (3.10)

holds for a ∈ (0, 1/3].

When b <∞, we simply choose cε = b, and so

dε → d0 =

∫ 1

0

1

sJ 0sb(1) ds

which by Proposition 3.4 is finite. In the case b = ∞, first choose a K > 0 such that

J 0a (1) ∈ [−3I0(1),−I0(1)] = [ − 3

32π ,− 132π ] for a ≥ K , which is possible thanks to

Proposition 3.4. Define the map, for c ≥ 1,

f (c) = c−∫ 1

Kc−1

1

sJ 0sc(1) ds.

Clearly f is continuous and f (c) ≥ c, in fact one has the bounds c + λ log c ≤ f (c) ≤c + λ−1 log c with some λ > 0 for large enough c. Therefore, there exists a function f

so that f (f (c)) = c for all sufficiently large c. We then set cε = f( |log ε|32π − bε). Clearly

cε → ∞, from cε ≤ |log ε|32π − bε we have εcε → 0, and

d0 = limε→0

∫ Kc−1ε

ε

1

sJ 0scε(1) ds = 0,

using Proposition 3.4 once more.

Proof of Proposition 3.4. For this proof we denote points in R4 as z = (t, x1, x2, x3). The

first two claims of the proposition are obvious, the third one requires some calculation.

Denote

Nt(x) =1t>0√πt

exp(

− x2

t

)

and note the identities

Nt(x) = λNλ2t(λx), Nt(x)Nt(y) = N2t(x+ y)Nt/2((x− y)/2).

We will also use the complementary error function Ercf(s) = 2∫∞s N1(x) dx. With these

notation one has the following identity for a ≥ 0:

∫ ∞

0

Ercf( 1√

t

)

Nt(a)t−1 dt =2 tan−1(a)

πa. (3.11)

Here we use the standard branch of tan−1, that is, tan−1(0) = 0, and so tan−1(a)a → 1 as

a→ 0. To see (3.11), denote π times the left-hand side by S(a). One then sees that

S′(a) =

∫ ∞

0

√π Ercf

( 1√t

)

t−1/2[a2

t2exp

(

− a2

t

)]

dt(

− 2a

a2

)

.

Page 24: Boundary renormalisation of SPDEs

24 Explicit boundary corrections

Since the quantity in [·] is a total derivative, one can integrate by parts and find the relation

S′(a) =2

a+ a3− 1

aS(a). (3.12)

One furthermore notices that

S(1) =

∫ ∞

0

1

2∂t

(

2

∫ ∞

1/√te−x

2

dx)2

dt =π

2. (3.13)

The differential equation (3.12)-(3.13) defines S uniquely and one can easily verify that

S(a) = 2 tan−1(a)a is a solution. This proves (3.11) and by scaling one gets

∫ ∞

0

Ercf( b√

t

)

Nt(a)t−1 dt =2 tan−1(a/b)

πa,

which also holds in the limiting case b = 0.

Note now that we can write

−J 0a/3(1) =

R4u

2K(1 − z)Ra(1, z) + 2K(1 − z0)Ra(1, z) − |Ra(1, z)|2 dz

= 4

∫ ∞

0

ae−arJ (−1, 1 + r) dr + 4

∫ ∞

0

ae−arJ (1, 1 + r) dr

− 4

∫ ∞

0

∫ ∞

0

a2e−a(r+r)J (1 + r, 1 + r) dr dr ,

(3.14)

where J is defined by

J (a, b) :=

R4u

N4t(x1)N4t(x2)N4t(x3 + a)N4t(x1)N4t(x2)N4t(x3 + b) dz .

For a, b such that a+ b ≥ 0, we then have

J (a, b) =

R4u

N8t(2x1)N8t(2x2)N8t(2x3 + a+ b)(N2t(0))2N2t(a− b) dz

=

R4u

1

23πtN2t(x1)N2t(x2)N8t(2x3 + a+ b)N2t(a− b) dz

=

t,x3>0

1

25πt3/2N1

(2x3 + a+ b√8t

)

Nt

(a− b√2

)

dx3 dt

=

t,x3>0

1

29/2πtN1

(

x3 +a+ b√

8t

)

Nt

(a− b√2

)

dx3 dt

=

t>0

1

211/2πtErcf

(a+ b√8t

)

Nt

(a− b√2

)

dt

=1

24π2

tan−1(

2a−ba+b

)

a− b,

(3.15)

and in particular J (1,−1) = 164π , which proves (3.6) as promised.

Page 25: Boundary renormalisation of SPDEs

Explicit boundary corrections 25

To show (3.10), we bound each integral appearing in (3.14) separately. To bound the

first two, notice that J (−1, 1 + r),J (1, 1 + r) . 1 ∧ r−1. One can then decompose the

integral as

a(

∫ 1

0

e−ar dr +∫ a−1

1

e−arr−1 dr +

∫ ∞

a−1

e−arr−1 dr)

.

Here the first integral is clearly bounded by 1, the second one by |log a|, and the third one

is independent of a, so it is also of order 1. Moving on to the last term in (3.14), we use the

the bound J (1 + r, 1 + r) . 1 ∧ (r + r)−1. One can then write

∫ ∞

0

∫ ∞

0

a2e−a(r+r)(1 ∧ (r + r)−1) dr dr =

∫ ∞

0

∫ r

0

a2e−ar(1 ∧ r−1) dr dr

= a2(

∫ 1

0

e−ar r dr +∫ ∞

1

e−ar dr)

.

Now the first integral is bounded by 1 and the second by a−1. This completes the proof.

3.3 Φ43 - the cubic term

For the Φ43 equation, there is one more term that is well below regularity −1, which is the

cube of Ψ. While it does not require additional boundary renormalisation, the fact that it

is ‘compatible’ with the boundary renormalisation of the square is far from obvious. This

‘compatibility’ is formulated in the following lemma, whose proof is the goal of this section.

Lemma 3.5. In the setting of Lemma 3.2, for any sufficiently small κ > 0 the sequencesδ∂R

1εΨε,cε, (RR2

ε)Ψε,cε, andR3εΨε,cε converge in C−3/2−κ(R4) to limits that do not depend

on ρ.

The first step is to show that the free field Ψ0,a can be restricted to the boundary ∂D –

even though it is only a distribution. This is not unlike the temporal restriction in [Hai14,

Sec. 9.4]. However, the present setting gets somewhat involved as we will need continuity

of this restriction not only with respect to the mollification, but also with respect to the

kernel as well as the hyperplane on which we restrict.

We start by introducing a few notations. Recall that R4u = (x0, x1, x2, x3) ∈ R

4 :x3 > 0. Take β ∈ (0, 5). Let Gβ be the set of functions G of the form

G =∑

n∈NGn, (3.16)

where Gn : R4 × R

4 → R such that suppGn ⊂ (x, y) : ‖x − y‖ . 2−n and

|DkxD

ℓyG

n(x, y)| . C2n(5+|k|+|ℓ|−β) for (x, y) ∈ (R4u∪−R

4u)2 with someC for all |k| ≤ K

and ℓ ≤ L. The optimal choice of C in the optimal choice of the decomposition (3.16)

yields a norm on Gβ . The values K and L will be occasionally relevant, in this case we use

the notation G (K,L)β .

For y ∈ R4u denote by y its projection to its first three coordinates. Let G ∂

β be the

set of functions G of the form (3.16) where this time Gn : R3 × R

4u → R such that

suppGn ⊂ (x, y) : ‖x − y‖ . 2−n, y3 ∈ In, where In ⊂ R are of size 2−n, and

|Gn| . C2n(5−β) with some C for all n ∈ N. The optimal choice of C in the optimal

Page 26: Boundary renormalisation of SPDEs

26 Explicit boundary corrections

choice of the decomposition (3.16) yields a ‘norm’ on G ∂β 1. On both Gβ and G ∂

β we

denote by W η the multiplication by (|y3| ∧ 1)η. The relevant properties of these spaces are

summarised below.

Lemma 3.6. (i) Define G(ε) by

G(ε)(x, y) =

G(x, z)ρε(y − z) dz, x, y,∈ R4u.

Then for all η > 0, β′ ∈ (β − 1, β), ε > 0, and G ∈ G (1,1)β one has

‖W η(G−G(ε))‖G (1,0)

β′

. εη∧(β−β′)‖G‖G (1,1)β

.

(ii) Define G(r) by

G(r)(x, y) = G((x, r), y), x ∈ R3, y ∈ R

4u.

Then for all β′ ∈ (β − 1, β), r, r′ ≥ 0, and G ∈ G (1,0)β one has

‖G(r) −G(r′)‖G ∂β′

. |r − r′|β−β′‖G‖G (1,0)β

.

(iii) Define the random variables ΦG(ϕ) by

ΦG(ϕ) = ξ(

y 7→∫

R3

G(x, y)ϕ(x) dx)

, ϕ ∈ C∞(R3).

Then for all η ∈ (0, 1/2), β ∈ (0, 5/2), κ > 0 and G ∈ W−ηG ∂β there is a random

distribution ΦG such that for all test functions ϕ one has ΦG(ϕ) = ΦG(ϕ) almostsurely. Furthermore for all p > 0 and compact K ⊂ R

3 one has the bound

E‖ΦG‖pC−5/2+β−η−κ(K). ‖G‖p

W−ηG ∂β

.

Proof. In all of the proofs by homogeneity we may and will assume that the norms appearing

on the right-hand side equal to 1.

(i) Take ε ∈ (0, 1] and write the trivial bound

|∂kx(Gn −Gn(ε))(x, y)| ≤ 2n(5+|k|−β) (3.17)

for |k| ≤ 1. If ε ≥ |y3|, then this yields

|∂kx(Gn −Gn(ε))(x, y)| ≤ εη|y3|−η2n(5+|k|−β).

If ε < |y3|, then we can also write

|∂kx(Gn −Gn(ε))(x, y)| . ε supy′

|∂kx∇yGn(x, y′)| ≤ ε2n(5+|k|+1−β).

Interpolating between this and (3.17) gives the required bound.

1Note that G ∂β is not actually a vector space! Scalar multiplication however is well-defined and our ‘norm’

is positive and one-homogeneous.

Page 27: Boundary renormalisation of SPDEs

Explicit boundary corrections 27

(ii) The definitions immediately yield that for |α| ≤ 1 one has

|(G(r),n −G(r′),n)(x, y)| ≤ |r − r′|α supx′

|∂|α|x3 Gn(x′, y)| ≤ |r − r′||α|2n(5+α−β).

By interpolation it also holds for |α| = β−β′, giving the desired bounds onG(r)−G(r′). As

for the support, (G(r),n −G(r′),n)(x, y) is supported on (x, y) : ‖x− y‖ . 2−n, y3 ∈ Inwith In = [r − 2−n, r + 2−n] ∪ [r′ − 2−n, r′ + 2−n].

(iii) Let ϕλ be a test function on R3 on scale λ. It follows from Gaussianity and

Kolmogorov’s Hölder estimate (for negative exponents) that it suffices to show that

∥y 7→

R3

G(x, y)ϕλ(x) dx∥

L2(R4). λ−5/2+β−η . (3.18)

Set Qn = y : d(y, suppϕλ) . 2−n, y3 ∈ In and Qn = Qn \ ∪∞j=n+1Q

n. It is clear that

one has∣

Qn

|y3|−2η dy∣

∣.

2n(−5+2η) if 2−n ≥ λ,

λ42n(−1+2η) if 2−n < λ.(3.19)

We also claim that one has on Qn,

∣|y3|η

R3

G(x, y)ϕλ(x) dx∣

∣.

2n(5−β) if 2−n ≥ λ,

λ−(5−β) if 2−n < λ.(3.20)

To see the first case in (3.20), notice that in the sum

yη3∑

m∈N

R3

Gm(x, y)ϕλ(x) dx (3.21)

only terms with 2−m & 2−n contribute. For each of these terms we use supremum bound

on W ηGn, and recall that ϕλ is normalised in L1(R3). Therefore we get a bound of order

2−m&2−n

2m(5−β) . 2n(5−β)

as claimed. To see the second case in (3.20), the sum over the terms 2−m & λ yields the

required bound precisely as above. On the terms 2−m . λ we bound ϕλ by its supremum

and use that |suppGm(·, y)| . 2−4m, and therefore

2−m.λ

λ−42(5−β)m2−4m . λ−(5−β)

as claimed. From (3.19)-(3.20) we can write

∥y 7→

R3

G(x, y)ϕλ(x) dx∥

2

L2(R4)

.∑

n∈N

Qn

|y3|−2η dy∣

∣supy∈Qn

∣|y3|η

R3

G(x, y)ϕλ(x) dx∣

2

.∑

2−n≥λ2(5−2β+2η)n +

2−n<λ

λ−(6−2β)2n(−1+2η) . λ−5+2β−2η ,

which is precisely (3.18).

Page 28: Boundary renormalisation of SPDEs

28 Explicit boundary corrections

Notice that for ε > 0, one has

ΦG(r)3cε,(ε)

= Ψε,cε(·, r). (3.22)

For ε = 0, the right-hand side looks like the restriction of a distribution to a hyperplane,

which is in general not allowed. However, the left-hand side is perfectly meaningful thanks

to Lemma 3.6, so we can use (3.22) to define the restriction of Ψ0,b to the hyperplanes

R4u + (0, 0, 0, r).

Corollary 3.7. In the setting of Lemma 3.2, for all κ > 0 the functions r → Ψε,cε(·, r) andr → Ψε,∞(·, r) converge in Cκ([0, 1], C−1/2−6κ(R3)) in probability as ε→ 0.

Proof. We only provide the argument for the first case since the second is easier. It follows

from Lemma 2.15 that the kernels G3cε,(ε) converge in G (1,1)2−κ . By Lemma 3.6 (i) the con-

vergence also holds in W−κG (1,0)2−2κ. By Lemma 3.6 (ii)–(iii), the functions r 7→ ΦG(r)

3cε,(ε)

=

Ψε,cε(·, r) converge in Cκ([0, 1], Lp(Ω, C−1/2−5κ(Q1))). Kolmogorov’s continuity theorem

shows that if p is large enough, then the convergence holds inLp(Ω, C(5/6)κ([0, 1], C1/2−5κ(Q1)))which is as required, provided we substitute (5/6)κ → κ.

Now we have all we need for the proof of Lemma 3.5. Recall from Section 2.1 the

notation Q and Q∂ and denote Qr = (x0, x1, x2) : (x0, x1, x2, r) ∈ Q.

Proof of Lemma 3.5. It follows from Proposition 2.6 and Corollary 3.7 that the products

R1ε Tr∂ Ψε,cε converge in C−1/2−κ,−1/2−2κ

∂2(Q∂). By Proposition 2.2, this implies the con-

vergence in C−1/2−2κ(∂). It remains to use the basic fact that δ∂Cα(∂) continuously embeds

into Cα−1−κ(R4).

The last term is even easier: Proposition 2.6 and the convergence ofΨε,cε inC−1/2−κ(Rd)

implies the convergence ofR3εΨε,cε in C1/2−κ,−3/2−2κ

∂2(Q), which by Proposition 2.2 implies

the claim.

The statement concerning R2ε is more involved. First recall that R2

ε does not depend on

the variables x0, x1, x2. Take a test function ϕλ on R4 on scale λ. Our goal is then to show

∫ 1

0

R3

1y∈Qr

(

R2ε(r)(Ψε,cε(y, r)ϕλ(y, r) −Ψε,cε(y, 0)ϕλ(y, 0))

−R20(r)(Ψ0,b(y, r)ϕλ(y, r) −Ψ0,b(y, 0)ϕλ(y, 0))

)

dy dr∣

∣. o(1)λ−3/2−κ,

(3.23)

where o(1) → 0 in the ε → 0 limit. Note that the inner integral has to be understood in a

distributional sense. This understanding is justified by Corollary 3.7 and by the fact that

muliplying with 1Qr is a well-defined and continuous operation on C−1/2−6κ. First we treat

the easy case when the support of ϕλ is separated from Q∂ by at least λ. In this case the

integrand simplifies to

1y∈Qr(R2ε(r)(Ψε,cε(y, r) −Ψ0,b(y, r)) + (R2

ε(r) −R20(r))Ψ0,b(y, r))ϕλ(y, r)

and keep in mind that λ ≤ r. By Corollary 3.7 we have the bounds ‖Ψε,cε(·, r) −Ψ0,b(·, r)‖C−1/2−κ . o(1) (in ε) uniformly in r as well as ‖Ψ0,b(·, r)‖C−1/2−κ . 1. By

Lemma 3.2 (ii) we further have |R2ε(r)| . r−1−κ uniformly in ε, r as well as |R2

ε(r) −Rε0(r)| . o(1)r−1−κ uniformly in r. Finally, notice that ϕλ(·, r) can be seen as λ−1 times

Page 29: Boundary renormalisation of SPDEs

Regularity structures and models 29

a test function on scale λ on R3. Combining these bounds show that the left-hand side of

(3.23) is bounded by

∫ 1

0

o(1)λ−3/2−κr−1−κ1λ≤r dr ≤

∫ 1

0

o(1)λ−3/2−3κr−1+κ dr . o(1)λ−3/2−3κ.

This is the required bound with 3κ in place of κ, yielding the claim in the case when the

support of ϕλ is separated from Q∂ by at least λ.

In the alternative case the boundary terms in (3.23) have to be taken into account. Recall

the elementary identity

a1b1 − a2b2 − a3b3 + a4b4 = (a1 − a2 − a3 + a4)b1 + (a3 − a4)(b1 − b3)

+ (a2 − a4)(b1 − b2) + a4(b1 − b2 − b3 + b4),

which holds for any "product" that is bilinear. In our situation the "product" will be

the action of C−1/2−κ(R3) on C1/2+κ(R3), and the terms will be a1 = Ψε,cε(·, r), a2 =Ψε,cε(·, 0), a3 = Ψ0,b(·, r), a4 = Ψ0,b(·, 0), and b1 = R2

ε(r)ϕλ(·, r), b2 = R2ε(r)ϕλ(·, 0),

b3 = R20(r)ϕλ(·, r), b4 = R2

0(r)ϕλ(·, 0). From Corollary 3.7 we have the uniform bounds

‖a1 − a2 − a3 + a4‖C−1/2−κ . rκ/6o(1),

‖a3 − a4‖C−1/2−κ . rκ/6,

‖a2 − a4‖C−1/2−κ . o(1),

‖a4‖C−1/2−κ . 1.

Furthermore, not only can ϕλ(·, r) be seen as λ−1 times a test function on scale λ on R3,

also ϕλ(·, r) − ϕλ(·, 0) can be seen as rκ/6λ−1−κ/6 times a test function on scale λ on R3,

for sufficiently small κ > 0. Hence using Lemma 3.2 with κ/12 in place of κ,

‖b1‖C1/2+κ . r−1−κ/12λ−3/2−κ,

‖b1 − b3‖C1/2+κ . r−1−κ/12o(1)λ−3/2−κ,

‖b1 − b2‖C1/2+κ . r−1−κ/12+κ/6λ−3/2−κ−κ/6,

‖b1 − b2 − b3 + b4‖C1/2+κ . r−1−κ/12+κ/12o(1)λ−3/2−κ−κ/6.

Since the exponent of r is greater than −1 in each of these terms, the outer integration in

(3.23) can be performed as before and and the claimed bound holds.

4 Regularity structures and models

4.1 General remarks

We will use a mild modification of the general black box theory of regularity structures.

The first tweak is a slight relaxation of the required bounds on models. Its formulation is

somewhat technical, but the moral of it is simply that it is sufficient to assume the bounds

of the correct order from the models on test functions supported away from the boundary

for any symbol of degree above the codimension of a given boundary.

Proposition 4.1. Let P be a boundary of codimension k and let T = (A,T,G) be aregularity structure. Assume that we are given mappings Π, Π, and Γ such that:

Page 30: Boundary renormalisation of SPDEs

30 Regularity structures and models

• Γ : Rd × Rd → G is continuous and satisfies Γxx = 1 and ΓxyΓyz = Γxz;

• For all x ∈ Rd, Πx maps T≤−k to S ′(Rd), while for all x ∈ R

d \ P , Πx maps T toS ′(Rd \ P ), such that as elements of S ′(Rd \ P ), Πxτ = Πxτ for all τ ∈ T≤−k;

• the identities Πy = ΠxΓxy and Πy = ΠxΓxy hold;

• on T≤−k, (Π,Γ) is a model with norm bounded by 1;

• the following bounds hold:

|(Πzτ )(ϕλz )| ≤ λα, |Γxyτ |β ≤ ‖x− y‖β−α,

for all α ∈ A, τ ∈ Tα with |τ |α = 1, z ∈ Rd \P , λ ∈ (0, 1] such that λ ≤ (1/2)|z|P ,

y ∈ Rd \ P such that ‖z − y‖ ≤ (1/2)|z|P , and β < α.

Then there exists a unique model of the form (Π,Γ) such that, as elements of S ′(Rd),Πxτ = Πxτ for all τ ∈ T≤−k and as elements of S ′(Rd \ P ), Πxτ = Πxτ for all τ ∈ T .Furthermore, the norm of (Π,Γ) is bounded by a constant depending only on T and P andthe map (Π, Π,Γ) 7→ (Π,Γ) is continuous in its natural topology.

Proof. The proof is virtually identical to that of [HP21, Thm C.5].

Remark 4.2. In this statement, the various ‘norms’ are taken over the entire space Rd, but

since the operation is local this can clearly be localised to compact regions.

The convergence of models in our setting do not directly follow from [CH16], but we

aim to minimise the additional arguments. Let us first very briefly summarise how the

convergence results are obtained in [Hai14, Sec 10], loosely following the notation therein.

To each basis symbol τ one associates functions W (ε;k)i τ (z;x; y1, . . . , yk) in k+2 variables.

Here ε ∈ [0, 1], k is a natural number, and i runs over some finite set. For any fixed z, by

Wiener’s isometry Ik, any such function yields a distribution in the variable x, living in the

k-th homogeneous Wiener chaos. The distribution Πεzτ is then defined as the sum of all

these random distributions over all the indices k, i. Let us point out here the the first slight

difference to [Hai14]: therein, due to the translation invariance, the dependence on z can

freely be ignored.

By [Hai14, Thm 10.7], the convergence of a sequence of models (Πε,Γε)ε∈[0,1] of the

above form follow from the bounds

E|(Π0zτ )(ϕλz )|2 . λ2|τ |+κ, (4.1)

E|(Π0zτ −Πεzτ )(ϕλz )|2 . ε2θλ2|τ |+κ, (4.2)

where κ, θ > 0 are arbitrary. By [Hai14, Prop 10.11], these bounds follow from

|〈W (0;k)i τ (z;x; ·),W (0;k)

i τ (z; x; ·)〉|.

ζ

(‖x− z‖+ ‖x− z‖)ζ‖x− x‖2|τ |+κ−ζ , (4.3)

|〈(W (0;k)i −W (ε;k)

i )τ (z;x; ·), (W (0;k)i −W (ε;k)

i )τ (z; x; ·)〉|. ε2θ

ζ

(‖x− z‖+ ‖x− z‖)ζ‖x− x‖2|τ |+κ−ζ , (4.4)

where the scalar product is understood in the k-fold tensor product of L2(Rd1) and the

sum ranges over a finite set of values of ζ ∈ [0, 2|τ | + κ + |s|). Moreover, due to

Page 31: Boundary renormalisation of SPDEs

Regularity structures and models 31

Proposition 4.1, whenever |τ | > −1, one only needs these bounds to hold for x, x with

|x|∂ . ‖x− z‖, |x|∂ . ‖x− z‖.

Unfortunately, even with this modification in mind, (4.4) will not always hold in our

setting. It will be complemented with the following criterion.

Proposition 4.3. Let z ∈ Rd, P ⊂ R

d a boundary, ε > 0, and W : Rd × Rkd1 → R be a

function that satisfies the bounds

|〈W(x, ·), W(x, ·)〉| . 1|x|P≤ε∑

ζ

(‖x− z‖+ ‖x− z‖)ζ‖x− x‖2|τ |+κ−ζ ,

with the sum as above. Then one has the bounds, with any sufficiently small θ > 0,

E

∣Ik

(

ϕλz (x)W(x, ·) dx)∣

2

. ε2θλ2|τ |+κ−2θ. (4.5)

Proof. First note that the λ . ε case is straightforward, even without using the indicator

function 1|x|P≤ε. Indeed, just as one proves (4.1) from (4.3), one can bound the left-hand

side of (4.5) by λ2|τ |+κ . ε2θλ2|τ |+κ−θ as required.

We therefore may assume ε . λ. The left-hand side of (4.5) can clearly be bounded by

∫ ∫

ϕλz (x)ϕλz (x)〈W(x, ·), W (x, ·)〉 dx dx∣

. λ−2|s|∑

ζ

∫ ∫

1|x|P≤ε,‖x−z‖≤λ,‖x−z‖≤λ(‖x− z‖+ ‖x− z‖)ζ‖x− x‖2|τ |+κ−ζ dx dx

. λ−2|s|∑

ζ

λζ∫ ∫

1|x|P≤ε,‖x−z‖≤λ,‖x−z‖≤λ‖x− x‖2|τ |+κ−ζ dx dx

. λ−2|s|∑

ζ

λζ∫ ∫

1|x|P≤ε,‖x−z‖≤λ,‖x‖≤λ‖x‖2|τ |+κ−ζ dx dx.

By the assumptions on ζ , the integral in x is finite and bounded by λ2|τ |+κ−ζ+|s|. The

integral in x is trivially bounded by ελ|s|−1, and hence we get the desired bound.

Applying the proposition with W = δW (ε;k)i τ (z; ·; ·), we see that if instead of (4.4) one

has the bounds

|〈(W (0;k)i −W (ε;k)

i )τ (z;x; ·), (W (0;k)i −W (ε;k)

i )τ (z; x; ·)〉|. 1|x|P≤ε

ζ

(‖x− z‖+ ‖x− z‖)ζ‖x− x‖2|τ |+κ−ζ , (4.6)

then (4.2) holds.

Remark 4.4. Clearly the convergence criteria (4.3), (4.4) and (4.6) have the property that

if they hold for W (ε;k)i τ , then they also hold for 1x∈AW (ε;k)

i τ , for any measurable set A.

Informally speaking, the multiplication of Gaussian models with indicator functions is

straightforward.

Page 32: Boundary renormalisation of SPDEs

32 Regularity structures and models

Next, we need some analogues of [Hai14, Lem 10.14] in the case where the blowup of

a kernel is not only controlled by the distance to the diagonal but also to the boundary. For

α, γ ≤ 0, denote by ‖K‖α,γ the best proportionality constant in the bound

|K(x, y)| . (|x|∂ + ‖x− y‖)α‖x− y‖γ . (4.7)

When α = 0, then ‖ · ‖0,γ coincides with ‖ · ‖γ;0 from [Hai14, Sec 10.3], with the latter

0 indicating that no derivatives are involved in the bounds. Note also the trivial property

that negative powers can be transferred from the first component to the second, that is,

‖K‖α+α′,γ−α′ ≤ ‖K‖α,γ for α′ ∈ [0,−α].

Example 4.5. Consider the 1 + 1-dimensional homogeneous Neumann heat kernel on the

positive half line as in Example 2.9 and let the two terms in (2.6) be denoted byK1 andK2.

Then ‖K1‖0,−1 . 1 and ‖K2‖−1,0 . 1.

Lemma 4.6. (i) One has the bounds

‖KK‖α+α,γ+γ ≤ ‖K‖α,γ‖K‖α,γ .(ii) Set K(x, z) =

K(x, y)K(y, z) dy. Then in case

γ + γ > −|s|, α+ γ > −|s|, α+ γ + γ < −|s|one has the bounds

‖K‖α+γ+γ+|s|,0 . ‖K‖α,γ‖K‖0,γ .(iii) In case

γ, γ > −|s|, γ + γ < −|s|, α+ γ < −|s|one has the bounds

‖|x|−α−γ−|s|∂ K‖0,γ . ‖K‖α,γ‖K‖0,γ .

(iv) In caseγ, γ > −|s|, γ + γ < −|s|,

one has the bounds‖K‖0,γ+γ+|s| . ‖K‖0,γ‖K‖0,γ .

Proof. Claim (i) is trivial. Claim (iv) follows from [Hai14, Lem 10.14]. Concerning (ii)

and (iii), we divide the integral defining K into separate regions. In the case 2|x|∂ ≤‖x − z‖, we set A1 = ‖x − y‖ ≤ |x|∂, A2 = |x|∂ ≤ ‖x − y‖ ≤ (1/2)‖x − z‖,

A3 = ‖z − y‖ ≤ (1/2)‖x − z‖, A4 = Rd \ (A1 ∪ A2 ∪ A3). In the case 2|x|∂ ≥

‖x − z‖, we set B1 = ‖x − y‖ ≤ (1/2)‖x − z‖, B2 = ‖z − y‖ ≤ (1/2)‖x − z‖,

B3 = ‖x− y‖ ≤ 4|x|∂ \ (B1 ∪B2), B4 = Rd \B3. A schematic picture of the different

regions is given below, where thick lines denote balls with radii of order |x|∂ and thin lines

denote balls with radii of order ‖x− z‖.

x z

A1

A2 A3

A4

x z

B1 B2

B3

B4

Page 33: Boundary renormalisation of SPDEs

Regularity structures and models 33

First, we have

A1

K(x, y)K(y, z) dy∣

∣. |x|α∂

A1

‖x− y‖γ‖y − z‖γ dy . |x|α+γ+|s|∂ ‖x− z‖γ ,

where we have used that γ > −|s| in both of (ii) and (iii). This bound is clearly the right

order for (iii), while for (ii) the condition α+γ > −|s| implies |x|α+γ+|s|∂ ≤ ‖x−z‖α+γ+|s|.

Next,∣

A2

K(x, y)K(y, z) dy∣

∣. ‖x− z‖γ

A2

‖x− y‖α+γ dy.

For (ii), the condition α + γ > −|s| implies that the integral is of order ‖x − z‖α+γ+|s|,

yielding the required bound. For (iii), the integral is of order |x|α+γ+|s|∂ , also as required.

Further,

A3

K(x, y)K(y, z) dy∣

∣. ‖x− z‖α+γ

A3

‖y − z‖γ dy . ‖x− z‖α+γ+γ+|s|,

using γ > −|s|. For (ii), this is as desired. For (iii) the condition α + γ < −|s| implies

‖x− z‖α+γ+|s| . |x|α+γ+|s|∂ , yielding also a bound of the correct order. Finally,

A4

K(x, y)K(y, z) dy∣

∣.

A4

‖x− y‖α+γ‖y − z‖γ dy . ‖x− z‖α+γ+γ+|s|,

where we have used that α+ γ + γ < −|s| in both of (ii) and (iii). This is the same bound

as in the case of A3, and so it is of the right order.

We now move on the case 2|x|∂ ≥ ‖x− z‖. First, we have

B1

K(x, y)K(y, z) dy∣

∣. |x|α∂

B1

‖x− y‖γ‖y − z‖γ dy . |x|α∂‖x− z‖γ+γ+|s|.

For (ii), one has ‖x − z‖γ+γ+|s| . |x|γ+γ+|s|∂ , giving the required bound. For (iii), we

bound ‖x − z‖γ+|s| . |x|γ+|s|∂ . The integral over B2 is treated in exactly the same way.

Further,

B3

K(x, y)K(y, z) dy∣

∣. |x|α∂

B3

‖x− y‖γ‖y − z‖γ dy . |x|α∂∫

B3

‖x− y‖γ+γ dy.

For (ii), the condition γ+ γ > −|s| implies that the integral is of order |x|γ+γ+|s|∂ , yielding

the required bound. For (iii), the integral is of order ‖x − z‖γ+γ+|s| . |x|γ+|s|∂ ‖x − z‖γ ,

also as required. Finally,

B4

K(x, y)K(y, z) dy∣

∣.

B4

‖x− y‖α+γ‖y − z‖γ dy . |x|α+γ+γ+|s|∂ ,

where we have used that α + γ + γ < −|s| in both of (ii) and (iii). For (ii), this is the

required bound. For (iii), one can write |x|γ∂ . ‖x − z‖γ . This exhausts all cases and

finishes the proof.

Page 34: Boundary renormalisation of SPDEs

34 Regularity structures and models

We will also use the following fact frequently used in the singular SPDE literature (see

e.g. [HS17, HM18]). It is a simple tool to switch between convergences with respect to

different parameters, which can be useful in situations where convergence of an auxiliary

approximation is much easier to show.

Proposition 4.7. Let (aε,δ)ε,δ∈[0,1] be a two-parameter family in a metric space (A, d).Suppose that aε,δ → aε,0 as δ → 0 uniformly in ε ∈ [0, 1], and that aε,δ → a0,δ as ε → 0for any δ ∈ (0, 1]. Then aε,0 → a0,0. A quantitative version also holds: if for someC1, C2, γ1, γ2 > 0, γ0 ∈ R one has d(aε,δ, aε,0) ≤ C1δ

−γ1 and d(aε,δ, a0,δ) ≤ C2ε−γ2δγ0 ,

then for some C3, γ3 > 0 one has d(aε,0, a0,0) ≤ C3ε−γ3 .

4.2 PAM

In the case of the (PAM), the model is time-homogeneous. By the general machinery

developed in [BHZ19], the system formally given by

∆Y = ξ, ∆v = v|∇Y |2 − 2∇v · ∇Y,

determines a regularity structure T0 = (T0, G0, A0), with |Ξ| = −3/2 − κ for some

sufficiently small κ > 0. We use common pictorial representations of some elements of T0by writing for Ξ, thin lines for I , thick red lines for the abstract gradient of I , and when

joining red lines at their root we understand the scalar product of the terms. For example,

=∑3

i=1 I(DiIΞ)2

We then take a new regularity structure T = (T,G,A) by adding three new symbols:

we set T = T0⊕〈 , , X〉. The homogeneities of the new symbols are the same as for

the corresponding ‘red’ symbols, namely | | = −1− 2κ, | | = −4κ, | X | = −2κ,

so A = A0. The group G is isomorphic to G0, the action of its elements on the new

symbols being uniquely determined by setting G = and requiring the product to be

regular. We will also use the notation F∂ = , , X.

Recall from the setup of Section 3.1 that K stands for the truncated Green’s function

of the 3-dimensional Poisson equation and satisfies [Hai14, Ass. 5.1]. One can then use

the results of [CH16] to build the BPHZ models (Πε,Γε)ε∈[0,1] for T0, which converge

in probability as ε → 0, are admissible with respect to K , and satisfy ΠεxΞ = ξε, where

ξε denotes a mollified stationary white noise defined on all of R3. One easily sees that

Πεx = |∇K ∗ ξε|2 − ℓε( ) with ℓε( ) satisfying (3.2). Let us furthermore recall the

notation G for the Green’s function of the 3-dimensional Neumann Green’s function on

D. As such, (x, y) 7→ G(x, y) is 0 outside D × D. As a convention, by ∇1G we mean

the function that equals to the gradient in the x direction of G on D × D and 0 outside

D×D (as opposed to the distributional derivative of x 7→ G(x, y), which has an additional

boundary term).

As mentioned above, the extension of the Γε component to T is automatic. Concerning

the extension of Πε, we set for ε > 0

(Πεx )(y) = |Ψε|2(y) − E|Ψε|2(y),

(Πεx )(y) =(

(Πεx )(y))(

(Πεx )(y))

− ℓε( ),

(Πεx X)(y) =(

(Πεx )(y))

(y − x).

Page 35: Boundary renormalisation of SPDEs

Regularity structures and models 35

As in [HP21, App C], we also consider the models (Π+,ε,Γε) and (Π−,ε,Γε) defined by

(Π+,εx τ )(y) = (Πεxτ )(y)1R×D(y) and (Π−,ε

x τ )(y) = (Πεxτ )(y)1R×Dc (y). These models are

of course not admissible, but will be very handy in the analysis of the reconstruction operator

in Section 5 below.

Lemma 4.8. The three sequences of models (Πε,Γε), (Π+,ε,Γε), and (Π−,ε,Γε), converge.

Proof. We start by establishing a priori bounds on the models (Πε ,Γε), including a candidate

limit model for ε = 0. As discussed above, we phrase everything in terms of the functions

W and borrow some some graphical notations commonplace in the singular SPDE literature.

Each function W is given as an integral, which we represent by a graph as follows. Each

vertex of the graph represents a variable, and each edge a function of the two variables

corresponding to the two vertices it connects. The node denotes the basepoint z; denotes

the running variable x; denote the variables y1, . . . , yk (on which Wiener’s isometry acts);

denotes an auxiliary variable that is integrated out. The edge denotes a factor K(u−v),

where u and v are the two variables corresponding to the the two endpoints; denotes

a factor ∇K(u − v); denotes a factor ρε(u − v); denotes a factor ∇1G(u, v). For

example, Πεz has only one component in its chaos decomposition, which is the image

under Wiener’s isometry of the function W (ε;2) given by

W (ε;2) (z, x, y1, y2) = Kε(x− y1)Kε(x− y2) = .

In particular, W (ε;2) does not depend on z. Similarly, for the first ‘new’ symbol we have

W (ε;2) = .

Note that in the ε = 0 case the edge together with its endpoint becomes an integration

against a Dirac-δ so it can be dropped. This then defines W (0;2) and W (0;2) .

Later on we will also use the shorthand = − for the recentering of the

kernel K. A lot of our simplification relies on the following property: if a given edge

appears without any recentering in a graph, then the bound obtained in [CH16] on that

graph only uses the (pointwise) upper bound on that kernel. For example, the bound (4.3)

on the function W (ε;2) follow exactly in the same way as in the translation invariant case

for W (ε;2) , since the kernels Gε and Kε both satisfy the required bound of order −2.

This also immediately implies the required bounds for the symbol X.

Concerning the last remaining ‘new’ symbol and its translation invariant counterpart

, we have

W (ε;4) = , W (ε;2) = 2 , W (ε;0) = −2 ,

W (ε;4) = , W (ε;2) = 2 , W (ε;0)1 = −2 ,

Page 36: Boundary renormalisation of SPDEs

36 Regularity structures and models

W (ε;0)2 = 2 − 2 . (4.8)

The same argument as above provides the bound (4.3) for W (ε;4) , W (ε;2) , andW (ε;0)1 .

However, for W (ε;0)2 there is no corresponding term in the translation invariant case and

one has to work a bit more. First of all, the ε = 0 meaning of (4.8) is not obvious, since

both terms on their own diverge in the ε → 0 limit. Introducing = − , we can

rewrite (4.8) as

W (ε;0)2 = 2 + 2 . (4.9)

We now argue that these terms are each well-defined in the ε → 0 limit and satisfy the

bounds (4.2). In this computation we simplify the notation by only keeping track of the

orders of the singularities as defined in (4.7) of (combination of) edges, which we denote

by writing the order on the given edge. For example, instead of we write 0,−1 ,

or instead of we write −2, 0 . The graphs are then manipulated by using the

rules in Lemma 4.6. Both terms in (4.9) can be bounded by using Lemma 4.6 (ii) and (iv)

in the first inequality and (i) in the second one:

0,−1

−2, 0

0,−2 0,−2

0,−2

. 0,−1

−1, 0 0,−1

. −1,−2 . |log |x|∂ |.

Keeping in mind that since | | > −1 , it suffices to consider the case |x|∂ . ‖x − z‖,

this verifies (4.3) (notice that since k = 0 in this case, the L2-scalar product simplifies to

product of real numbers).

We now turn to the proof of convergence, for which we use Proposition 4.7. The auxiliary

approximations W (ε,δ) (and from them the models (Πε,δ,Γε,δ)) are built as follows. The

translation invariant part is simply the BPHZ model associated to the system

Y = ρδ ∗ K ∗ ρε ∗ ξ, v = ρδ ∗ K ∗ (v|∇Y |2 − 2∇v · ∇Y ).

On ‘new’ symbols we proceed similarly as above, by simply replacing the appropriate red

edges in the trees describing the functions W (ε,δ) by green ones.

For any fixed δ, the functions W (ε,δ) are uniformly smooth over ε ∈ [0, 1] in the xvariable (in fact with any Ck norm bounded by a power of δ). As a consequence, the

condition aε,δ → a0,δ as ε → 0 of Proposition 4.7 is met. The other condition will be an

easy consequence of the already established a priori estimates and the bound

ρδ(x− u)(∇1G(u, v) −∇1G(x, v)) du∣

∣. δθ‖x− v‖−2−θ + 1|x|∂≤δ‖x− v‖−2

for any θ ∈ (0, 1). Recall that in the translation invariant case the analogous bounds take

the simpler form

ρδ(u− v)(∂k1 K(v,w) − ∂k1 K(u,w)) dv∣

∣. δθ‖x− w‖−1−|k|−θ

Page 37: Boundary renormalisation of SPDEs

Regularity structures and models 37

for any θ ∈ (0, 1) and |k| = 0, 1. Therefore, for any τ , k, and i the function W (ε,δ;k)i τ −

W (ε,0;k)i τ has the following form: it can be represented by a tree similar to W (0,0;k)

i τ , with

each edge admitting the same bound as the corresponding edge in W (0,0;k)i τ , and at least

one edge is further multiplied with either a factor δ−θ or 1|x|∂≤δ. Therefore a bound similar

of the form (4.4) or (4.6) is satisfied, and the convergence (Πε,δ,Γε,δ) → (Πε,0,Γε,0) as

δ → 0, uniformly over ε ∈ [0, 1], follows. The convergence of the models (Π+,ε,Γε), and

(Π−,ε,Γε) then immediately follows from Remark 4.4.

4.3 Φ43

We proceed similarly to the above. Take the regularity structure T0 = (T0, G0, A0) and

the corresponding sequence of renormalised models (Πε,Γε) built for the Φ43 equation in

[Hai14]. We again use the periodic models with fundamental domain (−2, 2)3 and assume

the integration kernel being K, the truncated heat kernel in 1 + 3 dimensions, satisfying

[Hai14, Ass. 5.1]. We then take a new regularity structure T = (T,G,A) by adding five

new symbols: we setT = T0⊕〈F∂〉withF∂ = , , , , X. The homogeneities

of the new symbols are given by | | = −1−2κ, | | = −1/2−5κ, | | = −1/2−3κ,

| | = −4κ, | X | = −2κ.

For ε > 0 we then extend (Πε,Γε) to a family of models (Πε,a,Γε), a ∈ [0,∞], by

setting Πε,aτ = Πετ for τ ∈ T0 and

(Πε,ax )(y) = Ψ2ε,a(y) − EΨ2

ε,a(y),

(Πε,ax )(y) =(

(Πεx )(y))(

(Πε,ax )(y))

− 3ℓε( )(Πεx )(y),

(Πε,ax )(y) = Ψ3ε,a(y) − 3(EΨ2

ε,a(y))Ψε,a,

(Πε,ax )(y) =(

(Πεx )(y))(

(Πε,ax )(y))

− ℓε( ),

(Πε,ax X)(y) =(

(Πε,ax )(y))

(y − x).

Here the functions Ψε,a are defined as in (1.5), using the Robin heat kernels G3a that are

discussed in Section 2.3. The extension of Γε to the new symbols is again straightforward.

We also define the models (Π+,ε,a,Γε), (Π−,ε,a,Γε) exactly as before.

Lemma 4.9. Take a sequence (aε)ε∈[0,1] ⊂ [0,∞] that is convergent in the natural topology.Then the three sequences of models (Πε,aε ,Γε), (Π+,ε,aε ,Γε), and (Π−,ε,aε ,Γε), converge.

Proof. Following the argument in the proof of Lemma 4.8, we only really need to establish a

priori bounds for components of Πε,aτ whose counterpart for the corresponding translation

invariant symbol is completely cancelled by the renormalisation procedure.

For the Φ43 equation there are altogether two such terms: the non-recentered parts of

the 0-th chaos component of and the first chaos component of . For the first one, the

computation is very similar to the previous proof, and we use the graphical notation from

therein, but this time denoting a factor K(u − v) and denoting a factor G3aε(u, v).

The function W (ε;0)2 is then given by

W (ε;0)2 = 2 + 2 . (4.10)

Page 38: Boundary renormalisation of SPDEs

38 Proofs of the main results

The power counting also changes as now we have |s| = 5 and now admits the

bound −3, 0 . Consequently, we can bound both terms in (4.10) using Lemma 4.6 (ii)

and (iv) in the first inequality and (i) in the second one:

0,−3

−3, 0

0,−3 0,−3

0,−3

. 0,−3

−1, 0 0,−1

. −1,−4 . |log |x|∂ |.

Concerning the non-recentered term in the first chaos component of , it can be written as

W (ε;1)2 = 6 + 6 .

Therefore we have (using two instances of to denote the two different running variables

x and x) by applying Lemma 4.6 (ii) and (iv) in the second inequality, (i) in the third one,

and (iii) in the fourth and fifth one

|〈W (ε;1)0 (x; ·),W (ε;1)

0 (x; ·)〉| . 0,−3

−3, 0

0,−3 0,−3

0,−3

0,−3 0,−3

0,−3

−3, 0

0,−3 0,−3

0,−3

(4.11)

. 0,−3

−1, 0 0,−1

0,−1

0,−3

−1, 0 0,−1

. −1 + η,−4

0,−1

−1 + η,−4

. |x|η∂ 0,−1 −1 + η,−4

. |x|η∂ |x|η∂

0,−1

= |x|η∂ |x|η∂‖x− x‖−1 .

Here η < 0 is arbitrary and introduced only to avoid equalities in the conditions of

Lemma 4.6. Since |x|∂ . ‖x− z‖, |x|∂ . ‖x− z‖ can be assumed, this proves (4.3). The

remaining arguments are then exactly as in the proof of Lemma 4.8.

5 Proofs of the main results

The convergence results are obtained via [GH19, Thm. 5.6] (or a slight modification of it),

the notation of which we shall use without repeating them here. Let us however summarise

the step which is the main difference to most of the literature and the source of the boundary

renormalisation. This concerns setting up the convolution operator of a fixed point problem.

Given a modelled distribution f – think of the right-hand side of an abstract equation –

we will define an abstract convolution Pε(f ), where the operator Pε is built from the

following ingredients:

Page 39: Boundary renormalisation of SPDEs

Proofs of the main results 39

• A kernel K that is 2-smoothing in the sense of [Hai14, Ass. 5.1] – this will always

be the truncated (translation invariant) heat kernel for us;

• A model (Πε,Γε) admissible for K – we will always choose from the models dis-

cussed above;

• A remainder kernel Zε that is smooth locally, but not necessarily globally up to the

boundary;

• A distribution ζ that agrees with Rεf away from the spatial boundary R× ∂D.

Given these ingredients, [GH19, Lem 4.12 & 4.16] yields a modelled distribution Pε(f )

that satisfies

Rε(Pε(f ))(x) = 〈(K + Zε)(x, ·), ζ〉.

Moreover, Pε is locally Lipschitz continuous in the natural topologies. In particular, to

get convergence results, one has to ensure that the choice of ζ is sufficiently stable in the

ε → 0 limit. This is far from obvious: for the type of weighted spaces we wish to use,

the reconstruction theorem [Hai14, Thm 3.10] can only be applied locally in the interior of

R × D, which therefore yields a distribution Rεf only on the interior. Extending it to a

distribution ζ on the whole space, may not be obvious or indeed even unique. Of course

as long as ε > 0, there is trivial canonical extension, but it is not true in general that this

choice will converge to a limit. To ensure convergence, we modify the canonical choice by

certain Dirac masses on the boundary, which in turn has the effect of changing the boundary

conditions (see Section 1.3 or [HP21, Rem 1.5]). In line with [GH19], we will also use

the notation ζ = Rεf , to emphasise its relation to the ‘usual’ reconstruction operator from

[Hai14].

Let us recall that in the weighted spaces of modelled distribution used in [GH19]

weights are given by triples w = (η, σ, µ) ∈ R3 describing singularities near the temporal

and spatial boundaries and their intersection, respectively. It will often be convenient to use

the notation (η)3 = (η, η, η). The lowest homogeneity symbol appearing in the target space

of a given space of modelled distributions will usually be denoted by α. There are two

general situations when the issue of choosing ζ can be avoided. First, as long as σ∧α > −1,

there is a canonical choice for R, see [GH19, Thm 4.9], which furthermore agrees with the

natural extension for smooth models. Second, if σ ∧ α > −2 and K + Zε is a Dirichlet

kernel, then Pε(f ) can be defined without defining Rf , see Lemma A.3 below.

Finally, we address the above mentioned ‘slight modification’ of [GH19, Thm 5.6] that

we will use in the case of Φ43, namely that we do not consider the remainder part of the kernel

fixed as ε varies. It is straightforward to check that the integration operation against smooth

remainders with singularities at the boundary [GH19, Lem 4.16] is Lipschitz continuous

with respect to the natural norm ‖ · ‖Zβ,Pon the space of remainder kernels. One then gets

the following simple extension of [GH19, Thm. 5.6].

Proposition 5.1. Assume the setting of [GH19, Thm. 5.6], with however allowing forZ ∈ Zβ,P to depend on ε ∈ [0, 1]. Suppose that

‖Zε − Z0‖Zβ,P→ 0.

Then the conclusions of [GH19, Thm. 5.6] hold.

Page 40: Boundary renormalisation of SPDEs

40 Proofs of the main results

5.1 PAM

As a warm-up example of the above setup without boundary renormalisation, define the

modelled distribution

Yε = −Pε(1D ),

withZε = Z being the Neumann remainder (see either [GH19, Ex 4.15], [HP21, App A], or

Section 2.3 above) and ζ = 1Dξε. Here and below 1D is the indicator function of D. Note

that one can easily make sense (probabilistically) of the product 1Dξ, and one moreover has

1Dξε → 1Dξ in C−3/2−κ. It is also easy to see that 1D ∈ Dκ,(−3/2−κ)3P . Therefore, setting

wε = DYε, we have that wε → w0 ∈ D1+κ,(−1/2−κ)3P and Rε

wε = Ψε, for all ε ∈ [0, 1].

This allows us to set up the abstract fixed point problem corresponding to (tPAM) as

follows. Let 1+D denote the indicator function of R+ × D. Let Zε = Z be the either

the Dirichlet remainder (for Theorem 1.1) or the Neumann remainder (for Theorem 1.2).

Let furthermore iε be the modelled distribution responsible for the contribution of the

initial condition: it is simply obtained by taking the Taylor polynomial lift of the function

t, x 7→ 〈(K + Z)t(x, ·), u0eYε〉. The abstract counterpart of (tPAM) then reads as

vε = Pε(1+Dvε|wε|2)− 2P ′

ε(1+DDvε · wε) + iε. (5.1)

The operators Pε and P ′ε will only differ in the choice of Rε. In fact as we will shortly

see, for the latter this choice will be trivial.

We now specify the spaces of modelled distributions on which (5.1) can be shown to

be a contraction and therefore to have a unique solution. Denote by T≥α0 the subspace

of T0 generated by basis vectors of homogeneity greater or equal to α (note that no new

symbols are included in these spaces) and by T the polynomial subspace. Fix the exponents

γ = 3/2 + 3κ, γ0 = κ, α = 0, α0 = −1− 2κ, α1 = −1/2 − 3κ, the sectors

V = I(T≥α0

0 ) + T , V0 = T≥α0

0 , V1 = T≥α1

0 ,

as well as

η = −1/2 + 3κ σ = 1− 3κ µ = −1/2 + 3κ;

η0 = −3/2 + κ, σ0 = −1− 2κ, µ0 = −3/2 + κ;

η1 = −2 + κ, σ1 = −1/2− 4κ, µ1 = −2 + 2κ.

We use the shorthand w = (η, σ, µ) and similarly for w1 and w2.

It is easy to verify that, by [GH19, Lem 5.4], iε → i0 in Dγ,w(V ), provided κ is

sufficiently small. It also follows from [GH19, Lem. 4.3] that the mappings

f 7→ f |wε|2, f 7→ Df · wε

are globally Lipschitz continuous fromDγ,w(V ) toDγ0,w0(V0) andDγ0,w1(V1), respectively.

Furthermore, since σ1 ∧ α1 > −1, [GH19, Thm 4.9] provides a continuous extension Rε

of the reconstruction operator on Dγ0,w1(V1), which agrees with the canonical extension for

ε > 0. This now completes the definition of P ′ε in (5.1).

Furthermore, in the case of Dirichlet boundary conditions, Pε is also well-defined,

thanks to Lemma A.3. By [GH19, Thm 5.6] one can conclude that, provided κ > 0 is

sufficiently small, it has a unique local in time solution vε. Furthermore, since the right-

hand side is globally Lipschitz continuous (in fact, linear), one can easily extend the local

solutions to the whole time interval [0, 1]. Now the proof can be easily concluded.

Page 41: Boundary renormalisation of SPDEs

Proofs of the main results 41

Proof of Theorem 1.1. Invoking the usual arguments concerning the renormalisation in the

bulk (see e.g. [Hai14, Sec 9]), one sees that the function vε := Rεvε solves (tPAM)

with 0 Dirichlet boundary conditions. Hence vεe−Yε solves solves (ε-PAM), trivially also

with 0 Dirichlet boundary conditions, and therefore it coincides with uDirε . By [GH19,

Thm. 5.6], vε → v0, the latter clearly does not depend on ρ, and so the limit limε→0 uDirε =

(R0v0)e−Y0 =: uDir is also independent of ρ, yielding the claim.

In the Neumann case, there is no easy way to construct (or avoid constructing, as above)

Rε for the integration operator Pε. The main remaining step is therefore to build Rε as a

continuous map from Vε := Dγ,w(V )|wε|2 to C−3/2+κ, which is where all previous steps

in the article will be used. First of all, notice that wε is of the form

|wε|2(x) = + 2ϕε(x) · + |ϕε(x)|2 (5.2)

for some (T )3-valued ϕε. One however can use the newly added symbols to out regularity

structure to write an equivalent function

|wε|2(x) := + ϕε(x), (5.3)

where ϕε(x) = E|Ψε|2(x) − ℓε( ). Although the latter expression seems ill-defined for

ε = 0, it is clear that ϕε has pointwise limits away from the boundary. In fact, in the

notation of Lemma 3.1 we have the expression ϕε(x) = R2ε + R3

ε . The | · |2 in (5.3)

is merely notational and serves only as an analogy with (5.2). One then trivially has

Rε|wε|2 = Rε|wε|2, in fact one has Rεf |wε|2 = Rεf |wε|2 for all f ∈ Dγ,w(V ). Let us

therefore define the isometry ιε : Vε → Vε := Dγ,w(V )|wε|2 by replacing the |wε|2 factor

with |wε|2. We can conclude that it suffices to construct Rε on Vε, and then it can be pulled

back to Vε by ιε.

Lemma 5.2. For all ε ∈ [0, 1] there exist maps Rε : Vε → C−3/2+κ with the followingproperties.

• For all ε ∈ [0, 1], g ∈ Vε, and ψ ∈ C∞0 (R+ ×D) one has

〈Rεg, ψ〉 = 〈Rεg, ψ〉; (5.4)

• The bound

|Rε(f ε|wε|2)− R0(f0|w0|2)|−3/2+κ; T . |||f ε; f0|||γ,w;T + o(1) (5.5)

holds as ε → 0, uniformly in modelled distributions f ε ∈ Dγ,w(V,Γε), f0 ∈Dγ,w(V,Γ0) bounded by a constant C , and in T ∈ (0, 1];

• For ε > 0, f ∈ Dγ,w(V ), one has the identity

Rε(f |wε|2) = Rε(f |wε|2)− (aρ +|log ε|8π )δ∂(Rεf ), (5.6)

where aρ is as in Lemma 3.1.

Proof. By the definition of the model and Lemma 3.1, for ε > 0 one has the identity

Rε(f |wε|2) = Rε(f ) + (Rεf )δ∂R1ε + (Rεf )RR2

ε + (Rεf )R3ε

Page 42: Boundary renormalisation of SPDEs

42 Proofs of the main results

+ (Rεf )(aρ +|log ε|8π )δ∂ . (5.7)

Therefore it suffices to show that the first four terms on the right-hand side extend in an

appropriately continuous way to ε = 0.

First note that if f ∈ Dγ,w(V ), then f ∈ Dγ′,w′

with γ′ = γ − 1 − 2κ = 1/2 + κand w′ = w + (−1− 2κ)3, that is, η′ = −3/2 + κ, σ′ = −5κ, µ′ = −3/2 + κ. Although

one has σ > −1, unfortunately α′ = −1− 2κ < 1, so there is no canonical reconstruction

on Dγ′,w′

as [GH19, Sec 4.3] does not apply. This barrier is exactly what is overcome

in [HP21, App C]: since in our models we have a decomposition Πεx = Π+,εx + Π−,ε

x , by

[HP21, Thm C.5], there exists maps Rε0 : Dγ′,w′ → C−3/2+κ satisfying (5.4) and the desired

continuity properties. Furthermore, for ε > 0 one has the identity Rε0(f ) = Rε(f ),

therefore indeed extending the first term on the right-hand side of (5.7).

For the other terms we may forget that f is a modelled distribution and consider

Rεf as a generic element F of C1−2κ,wP . By Proposition 2.8 (a) TrR+×∂D maps C1−2κ,w

P

continuously into Cσ,η∧µP0(R+ × ∂D) = C1−3κ,−1/2+3κ

P0(R+ × ∂D). By Proposition 2.8

(b) multiplication on the latter space with C1−κ,−κ∂2

(Q∂) is continuous in both arguments

to C1−3κ,(−1/2+3κ,−κ,−1/2+2κ)

P0,∂2(R+ × ∂D), which in turn is continuously embedded into

C−1/2+2κ(R× ∂D) by Proposition 2.8 (c). Finally, by Proposition 2.2, C−1/2+2κ(R× ∂D)

continuously embeds into C−3/2+2κ(R4). This handles the second term on the right-hand

side of (5.7).

Next, we claim that for F ∈ C1−2κ,wP (R+ × D), and R ∈ C1−κ,−1−κ

∂ (Q) such that

Dx1R = Dx2R = 0, the distribution (RR)F defined as

〈(RR)F,ϕ〉 =∫

R+×DR(z)

(

Fϕ(z) − Fϕ(π∂z))

dz

is meaningful for all ε ∈ [0, 1], belongs to C−3/2+2κ, and is a continuous function of both

arguments. First notice that away from the boundaries this is just a continuous function,

which furthermore has an integrable blowup at P0. In particular, the behavior on the set

z : |z|P0≤ 2|z|P1

is easily controlled as in [GH19, Sec. 4.3], and so we focus on test

functions centered on the region z : |z|P0≤ 2|z|P1

and supported away from P0. Take

R and F as above with norm 1, y ∈ z : |z|P0≤ 2|z|P1

and a normalised test function ϕ.

First consider the case λ ≤ |y|P1/2. Then we can simply write

|〈(RR)F,ϕλy 〉| = |〈RF,ϕλy 〉| ≤ supz∈suppϕλ

y

|RF (z)| . |y|−1−κP1

|y|−1/2+3κP0

≤ λ−3/2+2κ,

which is a bound of required order. In the case |y|P1≤ 2λ the boundary terms have to be

taken into account. One has

|〈(RR)F,ϕλy 〉| .∫

suppϕλy

|R(z)||F (z)(ϕ(z) − ϕ(π∂z)) + (F (z) − F (π∂z))ϕ(π∂z)| dz

.

suppϕλy

|z|−1−κP1

(|z|−1/2+3κP0

λ−6|z|P1+ |z|−3/2+6κ

P0|z|1−3κ

P1λ−5) dz

. |y|−1/2+3κP0

λ−1−κ + |y|−3/2+6κP0

λ−4κ . λ−3/2+2κ,

where we have used that since the exponent of |z|P1is above −1, the integral is finite.

Therefore, (RR)F behaves as required.

Page 43: Boundary renormalisation of SPDEs

Proofs of the main results 43

Finally, for the fourth term, it suffices to notice that C1−2κ,wP embeds continuously into

C−1/2+3κ(R4), on which by Proposition 2.6 multiplication with C1−κ,−1−κ∂2

(Q) is continuous

and maps into C−1/2,−3/2+2κ∂2

(Q), which in turn by Proposition 2.2 embeds continuously

into C−3/2+2κ(R4).

Using Rε from Lemma 5.2 in the operator Pε, the fixed point problem (5.1) is com-

pletely well-defined and one can use [GH19, Thm 5.6] to conclude that, provided κ > 0is sufficiently small, (5.1) has a unique local solution vε. Now the main statements follow

easily.

Proof of Theorem 1.2. Unlike in the Dirichlet case, the difference between Rε and the

canonical extension of Rε does effect the equation. Using (5.6), we have that the function

vε := Rεvε do not exactly satisfy (tPAM), but rather

(∂t −∆)vε = vε(|Ψε|2 − Cε) − 2∇vε ·Ψε − (aρ +|log ε|8π )δ∂vε,

with 0 Neumann boundary condition. That is however equivalent to (tPAM) with boundary

condition

∂nvε = −(aρ +|log ε|8π )vε.

Hence, vεe−Yε satisfies (ε-PAM) with boundary condition

∂n(vεe−Yε) = −(∂nYε)(vεe

−Yε) + (∂nvε)e−Yε = −(aρ +

|log ε|8π )vεe

−Yε ,

and so it coincides with uε. The proof is then finished as above.

5.2 Φ43

In the case of theΦ43, one has to keep in mind that the remainder kernel may change as ε→ 0.

As before, we start by defining the modelled distributions, for (ε, a) ∈ [0, 1] × [0,∞],

wε,a = Pε,a(1D ).

The reconstruction of 1Dξε is again straightforward. As for Zε, in the notation of Lemma

2.15, we choose it to be Z (3a). As before, it is straightforward that wε,a belongs to

D1+3κ,(−1/2−κ)3 , it depends continuously on (ε, a) ∈ [0, 1] × [0,∞], and that Rεwε,a =

Ψε,a.

Similarly to Section 3.3, one can see that Ψε,a can be restricted to the temporal hyper-

plane P0 = t = 0 (in fact this is already done in the case without boundaries in [Hai14]).

Let again be iε,a be the modelled distribution responsible for the contribution of the ini-

tial condition: it is simply obtained by taking the Taylor polynomial lift of the function

t, x 7→ 〈(K + Z (3a))t(x, ·), u0 − Ψε,a P0〉. Equation (1.6) then motivates to consider the

fixed point problems

vε,a = −Pε,a(1+D(v3ε,a+3v

2ε,awε,a))− 3P ′

ε,a(1+Dvε,aw

2ε,a)−P ′′

ε,a(1+Dw3ε,a)+ iε,a. (5.8)

The operators Pε,a,P′ε,a,P

′′ε,a are built from the models (Πε,a,Γε) and the remainder

Z (3a), the only difference in them will be the choice of Rε.

We now specify the spaces of modelled distributions on which (5.8) can be shown to

be a contraction and therefore to have a unique solution. Denote by T≥α0 the subspace

Page 44: Boundary renormalisation of SPDEs

44 Proofs of the main results

of T0 generated by basis vectors of homogeneity greater or equal to α (note that no new

symbols are included in these spaces). Fix the exponents γ = 1 + 3κ, γ0 = κ, α = 0,

α0 = −1/2− κ, α1 = −1− 2κ, α2 = −3/2− 3κ, the sectors

V = I(T≥α0

0 ) + T , V0 = T≥α0

0 , V1 = T≥α1

0 , V2 = T≥α2

0 ,

as well as

η = −2/3 + κ σ = 1/2 − 4κ µ = −2/3 + κ;

η0 = −2 + 3κ, σ0 = −1/2− κ, µ0 = −2 + κ;

η1 = −5/3 − κ, σ1 = −1− 2κ, µ1 = −5/3− κ;

η2 = −3/2 − 3κ, σ2 = −3/2− 3κ, µ2 = −3/2− 3κ.

It is easy to verify that, by [GH19, Lem 5.4], iε,a ∈ Dγ,w(V ) depends continuously on

(ε, a) ∈ [0, 1] × [0,∞], provided κ is sufficiently small. It also follows from [GH19,

Lem. 4.3] that the mappings

f 7→ f3 + 3fwε, f 7→ fw2ε

are locally Lipschitz continuous from Dγ,w(V ) to Dγ0,w0(V0) and Dγ0,w1(V1), respectively,

and that w3ε → w

30 in Dγ2,w2(V2). Furthermore, since σ0 ∧ α0 > −1, [GH19, Thm 4.9]

provides a continuous extension Rε of the reconstruction operator on Dγ0,w0(V0), which

agrees with the canonical extension for ε > 0. This now completes the definition of Pε in

(5.8).

Furthermore, in the case of Dirichlet boundary conditions, that is, a = ∞, we are in an

even better shape. Indeed, P ′ε,∞ and P ′′

ε,∞ are also well-defined, thanks to Lemma A.3.

By [GH19, Thm 5.6] one can conclude that, provided κ > 0 is sufficiently small, it has a

unique local solution vε,∞. Now the proof can be easily concluded.

Proof of Theorem 1.6. Invoking the usual arguments concerning the renormalisation in the

bulk (see e.g. [Hai14, Sec 9]), one sees that the function vε,∞ = Rεvε,∞ solves

(∂t −∆)v = −v3 − 3v2Ψε,∞ − 3v(Ψ2ε,∞ − Cε) − (Ψ3

ε,∞ − 3CεΨε,∞),

with 0 Dirichlet boundary conditions. Hence vε,∞ + Ψε,∞ solves solves (ε-Φ43), triv-

ially also with 0 Dirichlet boundary conditions, and therefore it coincides with uDirε . By

[GH19, Thm. 5.6], vε,∞ → v0,∞, the latter clearly does not depend on ρ, and so the limit

limε→0 uDirε = (R0

v0,∞) +Ψ0,∞ =: uDir is also independent of ρ, yielding the claim.

Remark 5.3. To see in which sense uDir vanishes on the boundary, first note that R0v0,∞ is

a continuous function with 0 trace. It is also clear that by definition, Ψε,∞ vanishes on the

boundary for ε > 0. By Corollary 3.7, Ψ0,∞ has a meaningful trace, which is also 0.

Now we move on to the setup of Theorem 1.8. First consider the case b ≥ 0. Take bεas in the theorem and cε as constructed from bε in Lemma 3.2. Consider the sequence vε,cε .

Since cε < ∞, we are out of the scope of Lemma A.3, so we have to construct Rε for the

operators P ′ε,cε and P ′′

ε,cε.

The former goes in essentially the same way as in the case of the PAM. We can

summarise the relevant properties of Rε in the following statement. Its proof follows from

the same arguments as that of Lemma 5.2 (simply using Lemmas 3.2 and 4.9 in place of

Lemmas 3.1 and 4.8, respectively) and is therefore omitted.

Page 45: Boundary renormalisation of SPDEs

Proofs of the main results 45

Lemma 5.4. Let aρ, bε, and cε as in Lemma 3.2. Then for all ε ∈ [0, 1] there exist mapsRε : Dγ,w(V )w2

ε,cε → C−1−2κ with the following properties.

• For all ε ∈ [0, 1] and ψ ∈ C∞0 (R+ ×D) one has

〈Rεg, ψ〉 = 〈Rεg, ψ〉; (5.9)

• The bound

|Rε(f εw2ε,cε) − R0(f0w

20,b)|w0;T . |||f ε; f0|||γ,w;T + o(1) (5.10)

holds as ε → 0, uniformly in modelled distributions f ε ∈ Dγ,w(V,Γε), f0 ∈Dγ,w(V,Γ0) bounded by a constant C , and in T ∈ (0, 1];

• For ε > 0, f ∈ Dγ,w, one has the identity

Rε(fw2ε,cε) = Rε(fw

2ε,cε) − (aρ + bε + cε)δ∂(Rεf ). (5.11)

Moving on to P ′′ε,cε, notice that for each ε ∈ [0, 1] there is only a single modelled

distribution, 1+Dw

3ε,cε, on which Rε needs to be constructed. Also notice that one has for

ε > 0Rε(1+Dw

3ε,cε) = 1

+D(Ψ

3ε,cε − 3ℓε( )Ψε,cε). (5.12)

Let us then set, for ε ∈ [0, 1],

Rε(1+Dw3ε,cε) = 1

+DΠ

ε,cε + 31+((δ∂R

1ε)Ψε,cε + (RR2

ε)Ψε,cε +R3εΨε,cε). (5.13)

First notice that by (5.12), (5.13), the definition of Πε,cε , and Lemma 3.2, we have that

for ε > 0Rε(1+Dw

3ε,cε) = Rε(1+Dw

3ε,cε) − 31

+(aρ + bε + cε)δ∂(Ψε,cε). (5.14)

In particular, Rε(1+Dw3ε,cε) and Rε(1+Dw

3ε,cε) coincide in the interior of R+×D, as required.

It is also clear that Rε has the required continuity properties: for the first term on the

right-hand side of (5.13) this follows from the convergence of the models, while for the

second this is precisely the content of Lemma 3.5 (including that the expressions in (5.13)

are actually meaningful when ε = 0).

Proof of Theorem 1.8. It follows from the above that the abstract equation (5.8) with cεin place of a admits local solutions vε,cε ∈ Dγ,w. Furthermore, by (5.11) and (5.14),

vε,cε = Rεvε,cε satisfy (1.6) with boundary conditions ∂nvε,cε = −3cεvε,cε. The argument

leading to (1.6) then shows that uε := vε,cε +Ψε,cε satisfy (ε-Φ43) with boundary conditions

∂nuε = 3(aρ + bε)uε and so it coincides with uε. By Proposition 5.1 we have that vε,cε →v0,b, the latter clearly does not depend on ρ, so the limit limε→0 uε = (R0

v0,b)+Ψ0,b =: ub

is also independent of ρ. Finally, one notices that u∞ coincides with uDir as defined in the

proof of Theorem 1.6.

The case b < 0 requires a very slight modification in order to reduce ourselves to the

case a = 0 in (1.4). Defining bε = bε− b, the sequence (bε)ε∈(0,1] satisfies (1.1) with b = 0.

Replacing the maps Rε from Lemma 5.4 and (5.13) by Rε defined as

Rε(fw2ε,cε) := Rε(fw

2ε,cε) + bδ∂(Rεf ),

Rε(1+Dw3ε,cε) := Rε(1+Dw

3ε,cε) + 31

+bδ∂Ψε,cε,

the proof is concluded just as above.

Page 46: Boundary renormalisation of SPDEs

46 Abstract Schauder estimate – the Dirichlet case

Appendix A Abstract Schauder estimate – the Dirichlet case

We wish to obtain a Schauder estimate in the special case of the Dirichlet heat kernel, but

without dealing with spatial extensions of Rf , even in the regime α ∧ σ ∈ (−2,−1). We

remark that a similar issue was encountered in [Lab19, Sec 3], but therein the problem

could be handled with a more ad hoc workaround. Let us first formulate what we mean by a

Dirichlet kernel. Simplify notations by setting, in the context of [GH19], P0 = 0×Rd−1,

P1 = R× ∂D, and

β = 2, m0 = 2, m1 = 1.

Denote furthermore Q = [−1, 1] ×D.

Assumption A.1. Let K (with a corresponding abstract integration map I) and Z be as

in [GH19]. Suppose furthermore that for all n ∈ N, Gn := Kn + Zn is non-anticipative,

symmetric, and vanishes on (Rd \ P1) × P1.

Note that this implies the following: for any y ∈ Q, Dk2Gn(x, y) = Dk

1Gn(y, x), and

the latter vanishes whenever x belongs to the boundary, so

supy

|Dk2Gn(x, y)| . |x|P1

supy,z

|Dk1∇2Gn(y, z)| . |x|P1

2n(|k|s+|s|−1). (A.1)

So, in the case |x|P1. 2−n, we improve the usual bound |Dk

2Gn| . 2n(|k|s+|s|−2) to

|Dk2Gn| . |x|P1

2n(|k|s+|s|−1).

Next, take a partition of unity (essentially the same as in the proof of [Hai14, Prop. 6.9])

with the following properties:

• One has the bounds |Dkϕm,i| . 2m|k|s , | suppϕm,i| . 2−m|s|, as well as the property

suppϕm,i ⊂ y : |y|P1∼ 2−m.

• ∑

m∈N∑ℓ(m)

i=1 ϕm,i(y) = 1 for all y ∈ Q, and there exists a constant C such that for

allm, i, the support of ϕm,i intersects with ϕm,j for at mostC indices j (in particular,

ℓ(m) . 2m(|s|−1)).

Let us set

Gm,in (x, y) = Gn(x, y)ϕm,i(y)

We shall only use these functions when 2−n ≥ |x|P1, in which case Gm,in is nonzero only

if n ≤ m. Also, for each n ≤ m there is a set An,m ⊂ 1, . . . , ℓ(m) with cardinality of

order 2(m−n)(|s|−1) such that Gm,in = 0 for i /∈ An,m. We can now define our candidate for

acting on G(x, ·) by distributions ζ on Q:

ζ(G(x, ·)) :=∑

2−n<|x|P1

ζ(Gn(x, ·)) +∑

2−n≥|x|P1

m∈N

i∈An,m

ζ(Gm,in (x, ·)), (A.2)

which is meaningful provided the infinite sums (the first one in n, the second one in m)

converge. Note also that if ζ happens to be a piecewise smooth function up to P1, vanishing

for negative times, then the following identity holds on Q+ = [0, 1] ×D:

ζ(G(x, ·)) =∫

QG(x, y)ζ(y) dy.

Page 47: Boundary renormalisation of SPDEs

Abstract Schauder estimate – the Dirichlet case 47

The improvement in the Dirichlet case will follow from the following simple property. It is

trivial from (A.1) that Gm,in (x, ·) is |x|P12n(|s|−1)2−m|s| times an approximate Dirac-delta,

that is, a function of the form ψ2−m

y , with some y ∈ Q and ψ ∈ B as usual. However,

thanks to the vanishing at the boundary, this can be improved as follows.

Proposition A.2. Suppose 2−n ≥ (2−m ∨ |x|P1). Define

Gm,in (x, ·) = 2m(|s|+1)|x|−1P1

2−n|s|Gm,in (x, ·).

Then for any x ∈ [−1, 1] ×D, the function Gm,in (x, ·) is of the form ψ2−m

y , with ψ ∈ B asusual, and |y|P1

≥ 2−m+1.

Proof. It suffices to show that (i) supp Gm,in (x, ·) ⊂ y : |y|P1∼ 2−m, (ii) | supp Gm,in (x, ·)| .

2−m|s|, (iii) |DkGm,in (x, ·)| . 2m(|s|+|k|s).

Property (i) and (ii) follows easily from the properties of ϕm,i. For property (iii), first

write

|DkGm,in (x, y)| . 2m(|s|+1)|x|−1P1

2−n|s|∑

k1+k2=k

sup|z|P1

∼2−m

|Dk1Gn(x, z)| supz

|Dk2ϕm,i(z)|

Since by assumption, is Gn = 0 on P1, one can write, using (A.1)

sup|z|P1

∼2−m

|Dk1Gn(x, z)| . 2−m sup|z|P1

.2−m

|∇Dk1Gn(x, z)| . 2−m|x|P12n(|s|+|k1|s)

where ∇ above denotes the spatial gradient. Along with the trivial bound |Dk2ϕm,i| .2m|k2|s , this yields (iii) as claimed.

To illustrate how this bound can be used, consider ζ ∈ Cα(Q) with α ∈ (−2,−1).

Extending it by 0 to R × D is of course not a problem, since the time boundaries have

codimension 2. One can write by (A.2) and Proposition A.2

|ζ(G(x, ·)| ≤∑

2−n<|x|P1

2−2n|ζ(22nGn(x, ·))|

+∑

2−n≥|x|P1

m∈N

i∈An,m

2−m(|s|+1)|x|P12n|s||ζ(Gm,in (x, ·))|

.∑

2−n<|x|P1

2−2n2−nα +∑

2−n≥|x|P1

|x|P12n|s|

m&n

2(m−n)(|s|−1)2−m(|s|+1)2−mα,

(A.3)

where to get the second inequality, we used that in the first line ζ is always applied to

approximate Dirac-deltas lying completely in R×D. Note that the infinite sums converge:

the exponent of 2n in the first and of 2m in the second, is −2− α < 0. Hence,

|ζ(G(x, ·)| . |x|2+αP1+

2−n≥|x|P1

2−n(2+α)2n|x|P1. |x|2+αP1

, (A.4)

where in the last inequality we used 1+α < 0. So indeed ζ(G(x, ·) is a well-defined function,

and in fact we even see the ‘improvement by order 2’ of the weight at the boundary, since

ζ(G(x, ·) vanishes on the boundary with speed 2 + α.

Page 48: Boundary renormalisation of SPDEs

48 Abstract Schauder estimate – the Dirichlet case

Fixing a γ > 0, we then define

Pf (x) = If (x) + J (x)f (x) + ZΠxf (x) + N f (x),

where we set

N f (x) =∑

|k|s<γ+2

Xk

k!(Rf −Πxf (x))(Dk

xG(x, ·)),

and the last term is understood in the sense of (A.2). We then have the following extension

of [GH19, Lem 4.12].

Lemma A.3. Suppose that Assumption A.1 holds. Fix γ > 0, w = (η, σ, µ), and let V bea sector of regularity α. Assume that these exponents satisfy

a∧ = η ∧ σ ∧ µ ∧ α ∈ (−2,−1),

and take

γ = γ + 2, η = (η ∧ α) + 2, σ = (σ ∧ α) + 2, µ ≤ (a∧ + 2) ∧ 0. (A.5)

Suppose furthermore that none of γ, η, σ, or µ are integers. Then, for f ∈ Dγ,wP (V ), one

has Pf ∈ Dγ,wP , where w = (η, σ, µ).

Furthermore, if (Π, Γ) is a second model realising K for I and f ∈ Dγ,wP (V, Γ), then

for any C > 0 the bound

|||Pf ; P f |||γ,w;K . |||f ; f |||γ,w;K + ‖Π− Π‖γ;K + ‖Γ− Γ‖γ;K

holds uniformly in models and modelled distributions both satisfying |||f |||γ,w;K+ ‖Π‖γ;K +

‖Γ‖γ;K + ‖ζ‖a,K ≤ C , where K denotes the 1-fattening of K.Finally, the identity

RPf (x) = Rf (G(x, ·)) (A.6)

holds on Q+, for f vanishing for negative times, where the right-hand side is understood inthe sense of (A.2).

Remark A.4. Recall the regularity of Rf , for f ∈ Dγ,wP (V ) as above: It is a distribution on

Q+ := [0, 1] ×D that satisfies the bounds

Rf (ψλx ) . λη∧α|x|µ−η∧αP1(A.7)

on Q+ ∩ |x|P0≤ 2|x|P1

for 4λ ≤ |x|P0, and

Rf (ψλx ) . λσ∧α|x|µ−σ∧αP0(A.8)

on Q+ ∩ |x|P1≤ |x|P0

for 4λ ≤ |x|P1. It then follows (see e.g. [GH19, Prop 2.15]) that

if η ∧ α > −2, then Rf has a unique extension as a distribution on Q such that it agrees

with Rf on Q+, it vanishes on [−1, 0] ×D, and it satisfies (A.7) on Q∩ |x|P0≤ 2|x|P1

for 2λ ≤ |x|P1

. In the sequel we denote this extension also by Rf . In particular, Rf ∈Cη∧σ∧µ∧α(Q).

Page 49: Boundary renormalisation of SPDEs

Abstract Schauder estimate – the Dirichlet case 49

Proof. Inspecting the proof of [GH19, Lem 4.12], there are two instances where an exten-

sion of Rf is required. On one hand, when establishing the bound (including making sense

of the left-hand side)

2−n≥|x|P0

|Rf (Dl1Gn(x, ·))| . |x|µ−|l|s

P1

( |x|P0

|x|P1

)(η−|l|s)∧0(A.9)

on the set |x|P0≤ |x|P1

, and the symmetric analogue on the set |x|P1≤ |x|P0

. On the other

hand, the extension of Rf is used in the estimate, for |k + l|s ≥ γ,

2−n≥|x|P0

‖x− y‖|k+l|s−γ |Rf (Dk+l1 Gn(y, ·))| . |x|η−γP0

|x|µ−ηP1(A.10)

whenever x, y, y satisfy 2‖x − y‖ ≤ 2‖x − y‖ ≤ |x|P0≤ |x|P1

, as well as the symmetric

analogue in the case |x|P1≤ |x|P0

. Our goal is therefore to make sense of these expressions

and derive the required bounds without extending Rf spatially, making use of Assumption

A.1. We note that, unlike in [GH19], the two cases |x|P0≶ |x|P1

have to be treated slightly

differently, since the special property of vanishing on the boundary is only assumed on P1.

We start with (A.9), when |x|P0≤ |x|P1

, first considering the terms |x|P0≤ 2−n ≤

|x|P1. By (A.7), we get

|Rf (Dl1Gn(x, ·))| ≤ 2−n(η∧α+2−|l|s)|x|µ−η∧αP1

≤ 2−n(η−|l|s)|x|µ−ηP1.

If η − |l|s > 0, then summing over |x|P0≤ 2−n ≤ |x|P1

gives the bound

|x|µ−|l|sP1

,

which is of the required order. For η − |l|s < 0 (equality cannot occur by assumption), the

summation yields

|x|η−|l|sP0

|x|µ−ηP1. |x|µ−|l|s

P1

( |x|P0

|x|P1

)η−|l|s. (A.11)

On the scale |x|P1≤ 2−n, we repeat the calculation in (A.3)-(A.4), making use of Rf ∈

Ca∧(Q) to obtain

2−n≥|x|P1

m∈N

i∈An,m

|Rf (Dl1G

m,in (x, ·))| (A.12)

.∑

2−n≥|x|P1

|x|P12n|s|

m&n

2(m−n)(|s|−1)2−m(|s|+1−|l|s)2−ma∧ . |x|2+a∧−|l|sP1

,

where in the summation with respect to n, we used that 1 + a∧ − |l|s is negative, so the

exponent of 2n is positive. Now suppose |x|P1≤ |x|P0

and consider the scale |x|P1≤

2−n ≤ |x|P0. Using again the argument in (A.3)-(A.4), this time making use of the

regularity of Rf given in (A.8), we get

|x|P0≥2−n≥|x|P1

m∈N

i∈An,m

|Rf (Dl1G

m,in (x, ·))|

Page 50: Boundary renormalisation of SPDEs

50 Abstract Schauder estimate – the Dirichlet case

.∑

|x|P0≥2−n≥|x|P1

|x|P12n

m&n

2−m(2−|l|s)2−m(σ∧α)|x|µ−σ∧αP0

. |x|µ−σP0|x|P1

|x|P0≥2−n≥|x|P1

2n(|l|s+1−σ).

If |l|s + 1 − σ > 0, then this gives a bound of order |x|µ−σP0|x|σ−|l|s

P1, which can be seen

to be of the right order just as in (A.11). If |l|s + 1 − σ < 0, then the above bound is of

order |x|µ−σ−1P0

|x|P1≤ |x|µ−σP0

, again as required. On the scales 2−n ≥ |x|P0, we can write

similarly to (A.12),

2−n≥|x|P1

m∈N

i∈An,m

|Rf (Dl1G

m,in (x, ·))| . |x|P1

|x|1+a∧−|l|sP0

≤ |x|2+a∧−|l|sP0

.

To prove (A.10), first recall that in the situation therein ‖x−y‖ . |x|Pi ∼ |y|Pi ∼ |y|Pi .Therefore by (A.9) one gets

2−n≥|x|P0

‖x− y‖|k+l|s−γ |Rf (Dk+l1 Gn(y, ·))|

. ‖x− y‖|k+l|s−γ |x|µ−ηP1|x|η−|k+l|s

P0. |x|η−γP0

|x|µ−ηP1,

where the condition |k + l|s > γ ≥ η was repeatedly used. The analogous estimate in the

case |x|P1≤ |x|P0

in this case follows exactly the same way.

References

[BBF18] I. Bailleul, F. Bernicot, and D. Frey. Space-times paraproducts for paracontrolled

calculus, 3d-PAM and multiplicative burgers equations. Annales scientifiques de l'Écolenormale supérieure 6, no. 51, (2018), 1399–1456. https://dx.doi.org/10.24033/

asens.2378.

[BGHZ19] Y. Bruned, F. Gabriel, M. Hairer, and L. Zambotti. Geometric stochastic heat

equations. ArXiv e-prints (2019). J. Am. Math. Soc, to appear. https://arxiv.org/

abs/1902.02884.

[BHZ19] Y. Bruned, M. Hairer, and L. Zambotti. Algebraic renormalisation of regularity

structures. Invent. Math. 215, no. 3, (2019), 1039–1156. https://arxiv.org/abs/

1610.08468. https://dx.doi.org/10.1007/s00222-018-0841-x.

[CH16] A. Chandra and M. Hairer. An analytic BPHZ theorem for regularity structures.

ArXiv e-prints (2016). https://arxiv.org/abs/1612.08138.

[CS18] I. Corwin and H. Shen. Open ASEP in the weakly asymmetric regime.Communicationson Pure and Applied Mathematics 71, no. 10, (2018), 2065–2128. https://dx.doi.

org/10.1002/cpa.21744.

[CvZ21] K. Chouk and W. van Zuijlen. Asymptotics of the eigenvalues of the Anderson

Hamiltonian with white noise potential in two dimensions. The Annals of Probability49, no. 4(2021). https://dx.doi.org/10.1214/20-aop1497.

[Ger20] M. Gerencsér. Nondivergence form quasilinear heat equations driven by space-time

white noise. Ann. Inst. H. Poincaré Anal. Non Linéaire 37, no. 3, (2020), 663–

682. https://arxiv.org/abs/1902.07635. https://dx.doi.org/10.1016/

j.anihpc.2020.01.003.

Page 51: Boundary renormalisation of SPDEs

Abstract Schauder estimate – the Dirichlet case 51

[GH19] M. Gerencsér and M. Hairer. Singular SPDEs in domains with boundaries. Proba-bility Theory and Related Fields 173, no. 3, (2019), 697–758. https://dx.doi.org/

10.1007/s00440-018-0841-1.

[GPS20] P. Gonçalves, N. Perkowski, and M. Simon. Derivation of the stochastic Burgers

equation with Dirichlet boundary conditions from the WASEP. Annales Henri Lebesgue3, (2020), 87–167. https://dx.doi.org/10.5802/ahl.28.

[Hai14] M. Hairer. A theory of regularity structures. Inventiones Mathematicae 198, no. 2,

(2014), 269–504. https://arxiv.org/abs/1303.5113. https://dx.doi.org/

10.1007/s00222-014-0505-4.

[HL15] M. Hairer and C. Labbé. A simple construction of the continuum parabolic Anderson

model on R2. Electron. Commun. Probab. 20, (2015), 11 pp. https://arxiv.org/

abs/1501.00692. https://dx.doi.org/10.1214/ECP.v20-4038.

[HM18] M. Hairer and K. Matetski. Discretisations of rough stochastic PDEs. The Annals ofProbability 46, no. 3(2018). https://dx.doi.org/10.1214/17-aop1212.

[HP15] M. Hairer and É. Pardoux. A Wong-Zakai theorem for stochastic PDEs. J. Math.Soc. Japan 67, no. 4, (2015), 1551–1604. https://arxiv.org/abs/1409.3138.

https://dx.doi.org/10.2969/jmsj/06741551.

[HP21] M. Hairer and É. Pardoux. Fluctuations around a homogenised semilinear random

PDE. Arch. Ration. Mech. Anal. 239, no. 1, (2021), 151–217. https://arxiv.org/

abs/1911.02865. https://dx.doi.org/10.1007/s00205-020-01574-8.

[HRW12] M. Hairer, M. Ryser, and H. Weber. Triviality of the 2D stochastic Allen-Cahn

equation. Electronic Journal of Probability 17, (2012), 1–14. https://dx.doi.org/

10.1214/ejp.v17-1731.

[HS17] M. Hairer and H. Shen. A central limit theorem for the KPZ equation. Ann. Probab.45, no. 6B, (2017), 4167–4221. https://arxiv.org/abs/1507.01237. https://

dx.doi.org/10.1214/16-AOP1162.

[Lab19] C. Labbé. The continuous Anderson hamiltonian in d ≤ 3. Journal of FunctionalAnalysis 277, no. 9, (2019), 3187 – 3235. https://dx.doi.org/10.1016/j.jfa.

2019.05.027.

[Par18] S. Parekh. The KPZ limit of ASEP with boundary. Communications in Math-ematical Physics 365, no. 2, (2018), 569–649. https://dx.doi.org/10.1007/

s00220-018-3258-x.