<<

arXiv:2004.09637v3 [math.PR] 21 Jan 2021 nnt iesoa SDEs dimensional Infinite 4 SDEs dimensional Finite 3 variables random Grassmann 2 Introduction 1 Contents model. Yukawa differen partial stochastic non-commutative probability, . nain esrs...... non-linearity . . . small . . . for . . behavior . . . Long-time . . . SDEs . of . measures 3.5 . solutions Invariant . for . formula drift . Itô 3.4 general An . for uniqueness . and 3.3 . Existence motion Ornstein–Uhlenbeck Grassmann 3.2 The 3.1 . rsmn asinvrals...... variables variables . Gaussian Grassmann Grassmann . on . and 2.3 . Topology . probability 2.2 Grassmann 2.1 Keywords: classification: subject A.M.S. uaamdl niaigas o hsmto a eextende be can method remo this the how of and also quantization indicating stochastic model, the Yukawa of th describe proof to a and give noise we Gra white Langevin space-time dimensional Gaussian infinite by proces driven and Gaussian finite Grassman of of in behavior notion values time the taking define equations We differential (partial) state. suitable a with eefo h on fve fqatmpoaiiy Grassma probab quantum a a into probability: algebra quantum Grassmann of abstract view an of of phism point Ana the theories. from field quantum here fermionic Euclidean of tization eitoueasohsi nlsso rsmn admvari random Grassmann of stochastic a introduce We rsmninsohsi nlssadtestochastic the and analysis stochastic Grassmannian rsmn lers ulda emo ed,sohsi q stochastic fields, Fermion Euclidean algebras, Grassmann uniaino ulda Fermions Euclidean of quantization rnec .D eci asmlaoGubinelli Massimiliano Vecchi, De C. Francesco [email protected],[email protected], [email protected], [email protected] [email protected] [email protected], asoffCne fMteais& of Center Hausdorff nttt fApidMathematics Applied of Institute egoAbvro ug Borasi, Luigi Albeverio, Sergio nvriyo on Germany Bonn, of University 12,8T8 60H15 81T08, 81S20, aur 2020 January Abstract 1 ileutos osrcieqatmfil theory, field quantum constructive equations, tial lers euete osuytelong the study to them use We algebras. n yi nGasanagba sdeveloped is algebras Grassmann on lysis i nain esrs sa application an As measures. invariant eir a ftesaecto o h Euclidean the for cut-off space the of val sansohsi ieeta equations differential stochastic ssmann e,Bona oinadstochastic and motion Brownian ses, lt pc,ie a i.e. space, ility be utbefrtesohsi quan- stochastic the for suitable ables oohrmdl fqatmfields. quantum of models other to d nrno aibei nhomomor- an is variable random nn 24 ...... 20 ...... 9 ...... atzto,non-commutative uantization, 18 ...... 21 ...... C 14 ...... ∗ 8 ...... agbaendowed -algebra 22 ...... 26 18 8 2 5 The Yukawa2 model 31 5.1 Osterwalder–Schraderfields ...... 32 5.2 Stochasticquantization ...... 34 5.3 Finite dimensional approximations in finite volume ...... 36 5.4 The infinite dimensional equation in finite volume ...... 38 5.5 Theinfinitevolumelimit...... 39

A Remarks on the Grassmann Gaussians 42 A.1 Gross-Palmerformulation ...... 42 A.2 A tracial formulation of Grassmann variables ...... 45 A.3 The real-wave representation of Grassmann Gaussians ...... 48

B Convergence of the perturbative : finite dimensionalcase 50

C Besov spaces of Banach algebras 52

D Convergence of the perturbative series: infinite dimensional case 54

E Other interacting models 60

1 Introduction

Euclidean quantum field theories (QFTs) [161, 65, 136] are a tool to construct and analyze mathematical models of relativistic quantum fields, that is the quantum theory of elementary particle in interaction satisfying the basic requirements of special relativity, i.e. Poincaré covariance and locality [82, 164, 95, 32, 79, 17, 15, 165, 58]. In the case of Bose–Einstein particles, the Euclidean theory is given by a probability measure on Schwartz distributions over the Euclidean space Rd. In other words, bosonic Euclidean fields ϕ are random distributions [167, 124, 27]. By a well–known result of Osterwalder and Schrader [65] (see also [181]), a basic set of properties, essentially Euclidean covariance, reflection positivity and regularity, is sufficient to reconstruct a well-behaved relativistic QFT from these probabilistic data. In the case of Fermi–Dirac particles (e.g. the electron), the Euclidean theory is expressed in terms of a linear functional over an infinite dimensional Grassmann algebra [150, 151, 123, 127, 128]. The value of this functional on Grassmann monomials gives the correlation functions (Schwinger functions) of the Euclidean theory, which are represented by Berezin [28, 29] of the type

¯ ¯ ¯ −SE(ψ,ψ) ¯ dψdψO(ψ, ψ)e hO(ψ, ψ)i = ¯ , (1) ¯ −SE (ψ,ψ) R dψdψe where the fields ψ, ψ¯ are the generators of the GrassmannR algebra, O is a functional of ψ and ψ¯ generating the Schwinger functions, and SE(ψ, ψ¯) is the Euclidean action function, usually the sum of a quadratic part and a polynomial interaction (possibly involving also bosonic fields, which then will have to be averaged according to an appropriate probability measure) [57, 52, 114]. The formula (1) is an algebraic expression, devoid of analytic meaning and is used as a convenient bookkeeping for various kinds of expansions used to define and study the properties of the Schwinger functions [136, 144, 114]. This is essentially due to the fact that Grassmann algebras do not have a natural notion of positivity, nor a norm, associated with them. For early discussions of related issues involving Euclidean fermions see e.g. [59, 91].

Stochastic quantization (SQ) [92, 47] is an approach to the construction of the correlation functions of an Euclidean QFT introduced by Parisi and Wu [131]. The basic idea is to consider an additional variable (usually a fictitious time) and interpret the Euclidean fields ϕ as the stationary solution to a stochastic partial differential equations (SPDEs) involving this additional variable and an external Gaussian source of noise ξ. This strategy has been employed to rigorously study bosonic Euclidean theories starting with the work by Jona-Lasinio and Mitter [93, 94, 30] on the stochastic quantization of the Φ4 model for a scalar particle in two (Euclidean) dimensions and with quartic interaction. More recently the work of Hairer on regularity structures [81] opened the way to the study of the three dimensional Φ4 model, see also [41, 103]. While the original implementation of stochastic quantization gives rise to parabolic SPDEs,

2 variants can be constructed involving elliptic [4, 3] or hyperbolic equations [77, 76]. See the introductions to [5] and to [73] for further references.

From physicists’ point of view, stochastic quantization gives an alternative approach to define and regularize quantum theories (especially theories with gauge invariance) [47, 30]. However, in the past few years, it has been realized that stochastic quantization has also interesting properties from the mathematical point of view. By solving the stochastic evolution it is possible to express the random fields ϕ of a (bosonic) Euclidean theory as well behaved functionals of the external noise, i.e. ϕ = ϕ(ξ), and this allows to study the non-perturbative features of the former by leveraging the Gaussian structure of the latter. Another conceptual advantage of stochastic quantization is that it is essentially insensitive to questions of absolute continuity with respect to a reference measure. On the other hand, lack of a suitable reference measure is the main difficulty of the Gibbsian point of view as expressed in equations of the form (1). This new perspective led to a series of results on the global space-time control of the stochastic dynamics [121, 74, 5, 119] and to a novel proof of the constructions of non-Gaussian bosonic Euclidean quantum field theories in three dimensions [73].

The main motivation of the present work is to develop a stochastic quantization of Euclidean fermionic quantum field theories (QFTs). In consideration of the Grassmannian nature of the fermionic Euclidean fields as expressed by (1), this must involve stochastic partial differential equations taking values in Grassmann algebras, including the study of their long time behavior, invariant “measures” and regularity properties. Additionally we aim for a theory which is both simple and effective and which can be used in the construction and analysis of Euclidean QFTs that involve fermions. The main differences between the bosonic and the fermionic cases is the need of extending the notion of stochastic processes to a non-commutative framework. The approach we follow is to frame this problem in the context of non-commutative probability (sometimes referred to as quantum or algebraic probabil- ity) [132, 31, 118, 163, 1], namely we define the relevant Grassmannian objects as non-commutative random variables. A non-commutative probability space (A,ω) is given by a C∗-algebra A and a state ω, which is a linear normalized positive functional on A. Inspired by the general approach of Accardi et al. [2] we will define random variables with values in a Grassmann algebra Λ as homomorphisms of Λ into A. We will not require the homomorphisms to respect the natural involution present in any C∗-algebra, since there is no canonical candidate for that in Grassmann algebras. It is useful to keep in mind how classical commutative random variables fit in this algebraic approach. Consider a X : Ω → M on a probability space (Ω, F, P), taking values in a manifold M and let B be the algebra of all measurable, complex bounded functions on M. All the relevant probabilistic information about X is encoded in the homomorphisms of algebras f ∈ B 7→ f(X) ∈ A where A is the algebra of bounded measurable complex random variables on the basic space Ω endowed with the linear functional given by the expectation E associated to the probability measure P defined on a σ-algebra F. In particular, this is the point of view with which SDEs on manifolds are defined [85].

The embedding of Λ into A allows to use the topology of A to do analysis on Grassmann algebras. Of course the associated analysis will not be canonical from the point of view of the Grassmann algebra itself, but it will turn out to be powerful enough to allow us to obtain a satisfactory theory of stochastic quantization for fermions. Let us mention another analogy which can help the reader to understand this point of view. The standard approach to study Gaussian processes in Hilbert spaces [67, 91, 83] requires to have a non-canonical embedding of the Hilbert space into a larger Banach space B which supports the Gaussian measure. This Banach space is not canonical and there are various possible choices for any given Hilbert space H. For example is associated with the Hilbert space H of functions γ on R+ with square integrable , but Wiener measure is supported on the Banach space C of γ Hölder functions for any γ < 1/2. In the case of random variables taking values in a Grassmann algebra Λ, the role of this “bigger” Banach space is played by a suitable Clifford sub-algebra of A which provides us with enough Grassmann generators to realize our homomorphisms and which has a canonical notion of norm. It will be for us just “a convenient place where to hang our (analytic) hat on”.

The main finding of our work is that, once one accepts this point of view, the rest of the analysis falls in place quite naturally. Objects like Grassmann , Brownian motion and free fields are

3 relatively simple to define and control. On top of them a novel theory of stochastic differential equations (SDEs) with values in Grassmann algebras can be initiated and carried to the point to be able to discuss their long time evolution and invariant “measures”. Moreover many considerations carry over to SDE in infinite dimensions and allow to define and solve, in particular, the non-commutative nonlinear stochastic partial differential equations (SPDEs) appearing in stochastic quantization, again in relatively simple terms. As a test of the effectiveness of our approach we give a proof of existence of the infinite volume limit of the massive Euclidean Yukawa model for small coupling [127, 128, 110] (in Appendix E it is shown how the results of this paper, in particular in Section 5, hold also for several other models of quantum fields involving fermions, including quantum electrodynamics in two dimensions with a positive photonic mass, pseudo-scalar Yukawa model, as well as the Gross–Neveu model in two dimensions and some simplified supersymmetric models ,including the Wess–Zumino model). At the end, we find that analysis of Grassmann SPDEs is in some respect much simpler than its commutative counterpart: the discussion of certain parts of the theory, like the existence of global space-time dynamics, are made relatively trivial by the boundedness of Grassmann Gaussian variables.

As we already remarked, our approach follows Osterwalder and Schrader [127, 128] who were the first to do rigorous analysis of Euclidean Fermions. See the review of Palmer [130] for a relatively recent account of their theory. They were using the language of Fock space operators and not of Berezin calculus, and this allowed them to consider rigorously expressions involving infinite dimensional Grassmann algebras. The Fock space approach has also been recently used to discuss in the context of Parisi– Sourlas dimensional reduction [101, 100] and Euclidean stochastic quantization [4, 3, 48]. The results of these papers could also be framed in the language of Grassmann random variables introduced in the present work. Our work provides also non-trivial examples of non-commutative non-linear SPDEs and of their qual- itative analysis. Analysis of partial differential equations in non-commutative algebras is not well devel- oped in general. Minkowski QFTs provide examples of PDEs in non-commutative algebras [17]. Other relevant examples we are aware of are the work of Rosenberg [142] on the (linear) Laplace equation on the non-commutative torus and the subsequent work on linear operators and functional spaces on non-commutative spaces (e.g. [145, 177]), works on the algebraic and geometric features in PDEs (see e.g. [135, 43]), the work of Khrennikov on linear equations and differential operators on superspace [99] and the work of Osipov [126] on solutions of quantum field equations via Wick kernels. However none of these works concern SPDEs and all rather use indirect methods to find solutions. We have left open the problem of removing the small-scale regularization in the stochastic quantization of the Yukawa model. In order to make progress in this direction one will have to understand the renormalization problem for such singular SPDEs in our setting, along the line of the recent work of Hairer and others [81, 75, 103]. Apart from the more delicate analysis required by the low regularity of the random fields, the removal of these will make appear renormalized Wick products of Euclidean fermion fields. These operators are unbounded in general and their appearance make more delicate to close the estimates, but we are confident that this can be achieved in our framework. It is our hope that this study of Grassmann stochastic equations could pave the way to a deeper understanding of non-commutative stochastic analysis and PDE theory in general.

Historical remarks.

The rest of this introduction (before we describe the structure of our work) is dedicated to discuss the connections of our work with previous research. The interest in studying time evolutions and stochastic processes with values in non commutative algebras arose in connection with the development of quantum mechanics in the 20s–30s of last century and later in quantum field theory, particularly around the 50s. In these theories one associates physical observables with operators evolving in time and acting on a, in general infinite dimensional, complex Hilbert space. Mathematically, this lead to the representation theory of certain operator algebras (CCR- algebras resp. CAR-algebras in the case of fermions). In this line, quantum fields require the study of certain infinite dimensional (non-commutative) al- gebras, the study of which has been particularly stimulated by developments in the period 1940–1960, where the mathematical description of systems with an infinite number of degrees of freedom, in partic- ular for quantum fields and statistical mechanics has been worked out. Let us mention the first works

4 on von Neumann algebras by Murray and von Neumann (see e.g. [174]) and the work on C∗-algebras developed originally and quite independently by I. M. Gelfand and his school (see, e.g., [62]), since the beginnings of the 40s, and by I. E. Segal and his school (see e.g. [153, 17]). Both types of algebras are as- sociative and equipped with an involution (the adjoint operation). Whereas C∗-algebras are looked upon as abstract normed algebras, without necessarily being realized as algebras of operators acting in Hilbert spaces, the study of von Neumann algebras (also called W ∗-algebras) rather involves looking at them as self-adjoint algebras of bounded operators in Hilbert spaces, closed in the weak operator topology. Both algebras can be looked upon as particular cases of normed algebras, that in the case of closedness are called Banach algebras. More precisely, C∗-algebras are normed algebras with a unit I and an involution a 7→ a∗, any element a of which satisfies kaa∗k = kak2. Although in principle every von Neumann al- gebra can be looked upon as a C∗-algebra, structural and conceptual developments in the study of von Neumann algebras involve particularly their realizations in Hilbert space. The mathematical literature on normed, Banach, C∗ and W ∗ algebras is very rich and there are a number of excellent monographs on them, see e.g., Naimark [122] (see also, e.g. [33]), and Zelazko [179] for normed spaces and algebras, and Kadison and Ringrose [97, 96] (and references therein) for the other type of algebras. For appli- cations in , in particular quantum mechanics, quantum field theory and statistical mechanics, there are monographs by Bratteli and Robinson [36, 37], Emch [54], Baumgärtel [23], as well as Baumgärtel and Wollenberg [24], Holevo [84], Segal [155], Schmüdgen [148]; for relativistic quantum fields, we refer particularly to the books by Araki [15], by Baez, Segal and Zhou [17] and Haag [79]. The articles by Summers [166] and Doplicher [18] survey particularly some recent developments related to application of such algebras in quantum field theory. Osterwalder and Schrader [127, 128] discovered that Euclidean fermions satisfy the relations defining a Grassmann algebra. However see the papers of Schwinger and Nakano for the first discussion of Wick rotation for fermions [150, 151] and also the papers [173, 176, 78, 107, 34]. Grassmann algebras constitute a class of associative non commutative algebras that were introduced by Grassmann in the second half of the 19th century. They are connected with Clifford algebras, that in turn are the underlying structure to the study of CCR and CAR algebras of quantum mechanics. All such algebras are nowadays part of the general theory of algebra as pioneered by E. Noether in the 20s-40s, see e.g. Van der Waerden [171, 172], Corry [45]. While the original paper of Osterwalder and Schrader [128] uses the same point of view which is developed here, the majority of the subsequent work with Euclidean fermions involved the formalism of Berezin’s [29]. This formalism has been subsequently developed into an analysis on Grassmann al- gebras, in the form of super-analysis [29]. For super-algebras and super-analysis see also, e.g., Rogers [141] and Leites [108], on a more heuristic level also DeWitt [49]. For a review see Pestov [133]. Berezin inte- grals have been useful in calculating fermionic systems, see e.g. Zinn-Justin [180], Izkyson and Zuber [88]. Such methods are also quite essential in most of the mathematical literature on fermionic quantum fields, including renormalization theory, see, e.g., Feldman, Knörrer and Trubowitz [56], Salmhofer [144]. Eu- clidean quantum field theories involving fermions have surprisingly simpler behavior than the bosonic ones and have been very much studied also from the rigorous point of view [61, 57, 110, 136, 52, 26, 114], mainly in the formalism of Berezin integration. In this context let us mention also the probabilistic rep- resentation introduced by de Angelis et al. [9] for finite-dimensional Grassmann-valued Markov processes using Poisson processes.

After Parisi and Wu seminal paper, the stochastic quantization of fermions has been discussed heuristi- cally in the physics literature, starting with Kakudo et al. [98], Fukai et al. [60], Damgaard and Tsokos [46] and She–Sheng and Ting–Chang [178], up to the recent paper of Efremov [53]. These authors stress particularly the relations with “Grassmann valued random variables”, solving formally partial differential equations with Grassmann valued Gaussian white noise. The invariant measures of physical interest are here described by averages with respect to Berezin integrals. However all these papers never properly discuss the analytical difficulties of considering (stochastic) differential equations in Grassmann algebras and the related stochastic analysis. From the point of view of , pioneering work in the theory of non-commutative processes has been done by Accardi, Streater and Hudson and Parthasarathy and others in the 70s– 80s, see e.g. [1, 132, 163, 31]. This includes the study of non-commutative Markov semigroups, non- commutative Brownian motion and stochastic analysis based on on it, in particular also SDEs but with no space variable dependence, hence no SPDEs. Let us also mention the work of Gross [72, 68, 69, 71]

5 applying Segal’s non-commutative integration theory [152, 125] to Clifford algebras. Work extending SDEs with values in algebras other than Grassmann algebras, like C∗ and von Neumann algebras are by authors like Streater and coworkers [21, 20, 19], Applebaum and Hudson [11, 10], Belavkin [25], Carlen and Kree [40], Gordina [66], Sinha and Goswami [162], Kümmer [102], see also the monographs by Meyer [118], Holevo [84], and the recent survey article by Cipriani [42] and the references therein. Grassmann variables are in a certain sense completely non-commutative numbers: as such they appear also in non-commutative central limit theorems [175] and in the setting of non-commutative processes with independent increments [149, 35]. A rigorous approach to infinite dimensional Grassmann algebras and the related stochastic analysis is that of Rogers [139, 137, 138, 140, 109] which essentially consists in looking at all the finite dimensional sub-algebras, use the Berezin integral to compute averages and require certain natural consistency conditions in order to obtain a projective system. A similar line of research was also carried on by Kupsch and Haba [104, 80]. These last works follows the observation of Hudson and Parthasarathy [86] and Le Jan [106] (see also Meyer’s book [118]) which developed a unified representation of the Bosonic and Fermionic Brownian motion (and related stochastic calculus) using Fock space techniques. None of these lines of research seems to have reached a stage where the theory is powerful enough to easily and naturally accommodate SPDEs (or even SDEs) of the kind needed in stochastic quantization. One of the technical problem involved is that (infinite dimensional) Grassmann algebras, as we already remarked, do not come with natural analytic structured suitable for non-linear analysis. Recent discussions of this problem can be found in works of Ivashchuk [89] and Alpay et al. [6].

More successful rigorous attempts to define Grassmann stochastic quantization were undertaken in a series of papers by Scherbakov et al. [160, 159, 158] using the locally convex topology obtained by considering all the possible correlation functions of the fields and the noise as the family of semi-norms. Convergence in this setting therefore corresponds to convergence of all the correlation functions. This approach has the disadvantage of not allowing the introduction of norms strong enough to study differ- ential equations via standard approaches. Despite this technical shortcoming, which makes all the proofs quite involuted, it should be said that these papers define and successfully solve equations with finitely many degrees of freedom (i.e. with small scale and large scale cut-offs) and then address the convergence when the large scale cut-off is removed (infinite volume limit) using cluster expansion techniques. Other results in this direction were obtained later on by Ignatyuk et al. [87], with applications in statistical mechanics.

The main aim of our work is to provide a framework suitable to discuss singular Grassmann-valued PDEs and in particular SPDEs. Besides proving methods for existence, uniqueness of local and global solutions, and of invariant measures, that can in principle be used for several applications, we go in details in one of the simplest but physically relevant models of quantum fields involving fermions. This is the Yukawa model, that couples a scalar (or pseudo-scalar) massive field with a pair of Fermi fields. Historically it was introduced by Yukawa in 1935 and further studied e.g. in [115]. Its mathematical construction was achieved, in space-time dimension 2, as part of the constructive approach to relativistic quantum fields, see e.g. [105, 116, 64, 50, 117, 111, 112, 55, 63]. The Euclidean approach was developed in [128, 156, 157, 70, 176]. Our work show that the Euclidean model can be obtained independently as a stationary state from a Grassmann-valued stochastic quantization equation. It is expected that also the high energy regularization can be treated in this framework.

Structure of the paper. In Section 2 we first define Grassmann random variables and present various related notions like compatibility and independence. A distance and its corresponding topology on a space of Grassmann-valued random variables is introduced in Section 2.1 and Section 2.2 and its completion is taken. Moreover, functions on the Grassmann algebra ΛV over a pre-Hilbert space V are introduced together with their right, resp. left, derivative as maps from ΛV to ΛV ⊗ V (resp. V ⊗ ΛV ), and a Taylor formula is proven. Then basic notions needed for formulating SDEs for Grassmann-valued processes are introduced. In particular, in Section 2.3 Grassmann Gaussian variables are discussed and constructed together with corresponding notions of Gaussian white noise and Brownian motions. Section 3 studies finite dimensional Grassmann differential equations with additive Gaussian white noise. We first study the linear case (Grassmann Ornstein–Uhlenbeck motion) and exhibit an explicit formula expressing its solution, in particular its stationary solution, that remains bounded when the linear drift is given by a strictly negative operator (Proposition 24). In Theorem 27 and Theorem 28 we

6 present a unique solution for the non-linear case given by an odd polynomial. Here we use an iteration of Picard’s type and fix-point arguments in a Banach space. Global existence is established, in our setting, essentially by means of a Gronwall inequality. An (average) Itô formula for solutions of finite dimensional Grassmann SDEs (Theorem 29) is proven in details. In Section 3.4 invariant measures are defined for the general case. For this we prove a criterium that can be looked upon as a Grassmannian analogue of the concept of infinitesimal invariance of measures (expressed via a Fokker–Planck equation) in the commutative case (Lemma 32). The existence of an invariant measure is proven in Theorem 33, in the case of total drift consisting of a linear term plus a non-linear odd-degree term. The long time behavior is studied in Section 3.5. Here the “lack of positivity” in Grassmann algebras is bypassed by assuming a smallness condition on the non-linearity and strict negativity of the linear part of the SDE. The main results here are Theorem 38, giving the convergence of an approximation to the unique stationary solution, and Theorem 39, that exhibits the invariant measure (state). Section 4 is devoted to extend the previous results to SDE in infinite dimensions of the kind relevant for stochastic quantization. This allows us to use these processes to represent the correlation functions of Euclidean fermionic theories like those described by (1). To achieve these results, the infinite dimensional setting is made precise, by taking as underlying space for the generalization of the Grassmann algebra, a suitable infinite dimensional Fréchet nuclear space V . The main problem to overcome is the choice of a suitable space G(V ) of homomorphisms for the Grassmann algebra over V to a C∗-algebra A. We present a choice in terms of operator-valued Besov–Hölder spaces Cs(Rd, A). The appropriateness of this setting, permitting to extend finite dimensional methods to the infinite dimensional case, is demonstrated by Theorem 53 and Theorem 54, that essentially show that smallness of the non-linear part and a generator of a strictly contractive semigroup on Cs(Rd) for the linear part, suffice for having a unique solution of the stochastic Grassmann infinite dimensional equation for all times, as well as its convergence to a stationary solution for long times. In Section 5, as an illustration of the use of our results for Grassmann SPDEs, we present an application to the study of the Yukawa model of Euclidean quantum fields. In Section 5.1 the Euclidean construction of the model, in two Euclidean dimensions and with ultraviolet (UV) and infrared (IR) cut-offs, is presented in our setting, following the construction in [128]. In Section 5.2 the corresponding stochastic quantization equation (SQE) is formulated and studied. Two main theorems are formulated and proved here. In the first one (Theorem 58) we obtain the representation of the relevant correlation functions for solutions of the SQE as averages with respect to the probability state, for small values of the coupling constant λ in the non-linear part of the equation. The second theorem, Theorem 59, achieves the removal of the space cut-off (infrared, or infinite volume limit, required for the model to be euclidean invariant). The proof is quite technical and on its way several results are established that have an interest also for other applications. In Section 5.3 finite dimensional approximations are constructed, involving the introduction of the model first on a lattice approximation of a finite torus and then passing to the full torus. The passage to the full space is obtained by introducing weighted Besov spaces. The convergence result is presented in Theorem 69. This then yields, in particular, a construction of the fermionic sector of the massive Yukawa Euclidean quantum field model in two dimensions, in the weak coupling region and with an ultraviolet cutoff, by our method of Grassmann-valued SPDEs. In Appendix A we briefly describe the tracial representation of Euclidean Fermi fields of Gross [70] and Palmer [130], its relation with our description of Gaussian Grassmann random variables (see Section 2.3) and another formulation based on Segal’s “real wave” representation [17]. In Appendix B we include some side results about the convergence of series expansions for the solutions to finite dimensional Grassmann SDEs, while In Appendix C we collect various technical results about functional spaces of Banach-valued functions. Appendix D contains the proof of convergence of perturbative series in the infinite dimensional case of Yukawa2 model and, finally, in Appendix E we review some further quantum field theory models that can be approached using the ideas presented in the present paper.

Acknowledgements. This work is supported by DFG via the grant AL 214/50-1 “Invariant measures for SPDEs and Asymptotics”. The authors are funded by the DFG under Germany’s Excellence Strategy - GZ 2047/1, project-id 390685813. M. G. would like to thank Hao Shen for comments on an earlier version of the paper.

7 2 Grassmann random variables

We denote by L(A, B) the space of linear maps between vector spaces A and B, if both A, B have topologies we consider all the maps to be continuous. We let L(A) = L(A, A). With Hom(A, B) we denote homomorphisms between algebras A and B.

2.1 Grassmann probability Algebraic probability. We consider a complex Hilbert space H (to be fixed later) and denote by A = L(H) the C∗-algebra of bounded operators with the operator norm. Moreover we assume to have a state ω on A, that is a positive definite functional normalized by the condition that ω(I)=1, where I is the identity in L(H). The pair (A,ω) is a (non-commutative or algebraic) probability space. We do not require the state ω to be either faithful, or tracial.

Grassmann algebras. Let V be a (finite dimensional or infinite dimensional and separable) real Hilbert space. Denote by ΛV the Grassmann algebra generated by V , i.e. the obtained by ⊗n quotienting the algebra T (V ) = ⊕n>0V by the two-sided ideal generated by elements of the form x ⊗ x for x ∈ V . We denote the product of two elements f,g ∈ ΛV by f ∧ g (or if there is no confusion with other products simply by fg).

Let (vα)α∈A be a fixed basis of V . As a consequence ΛV is spanned by elements of the form vA := vα1 ∧···∧ vαn and A = (α1,...,αn) is a n-tuple of elements of A with the convention that v∅ = 1 is the unit element in ΛV . When it is clear from the context we will denote the product in ΛV simply by vA = vα1 ...vαn . We recall that ΛV is Z2 graded in the sense that it splits into odd and even parts ΛV = ΛoddV ⊕ ΛevenV . On ΛV there is an super Hopf algebra structure with coproduct ∆:ΛV → ΛV ⊗ ΛV where ΛV ⊗ ΛV is the Z2-graded tensor product algebra which satisfies (f ⊗ g)(h ⊗ k) = (−1)|g||h|fh ⊗ gk, f,g,h,k ∈ ΛV, with | · | :ΛV →{0, 1} the even/odd grading. The coproduct ∆ is the algebra homomorphism such that ∆v =1 ⊗ v + v ⊗ 1 for all v ∈ V ⊆ ΛV and the counit ε :ΛV → C given by ε(vA)= IA=∅.

Random variables.

Definition 1 A V -Grassmann random variable Ψ is an algebra homomorphism from the Grassmann algebra ΛV into A. We denote by G(V ) = Hom(ΛV, A) the set of all such homomorphisms. We call the law of Ψ ∈ G(V ) the family of its moments ωΨ(F ) := ω(Ψ(F )) for all F ∈ ΛV , also represented by the linear functional ωΨ :ΛV → R. Note that Ψ ∈ G(V ) cannot be assumed to be a ∗-algebra-homomorphism since ΛV has no (natural)

∗-operation. If F ∈ ΛV has representation F = A FAvA we shall employ the following dual notation A F (Ψ) := Ψ(PF )= FAΨ , XA A where Ψ := Ψ(vA) and FA ∈ R. Since Ψ is assumed to be an algebra homomorphism, we have e.g. α β β α α Ψ Ψ = −Ψ Ψ , where Ψ = Ψ(vα) and (vα)α is a fixed basis of V . Moreover, Ψ(ΛV ) is a Grassmann sub-algebra of A and Ψ(F ) has the same degree (even or odd) as F . As shown in [171, 172], even if some arguments are formulated in a basis dependent way, the definition of ΛV and its characterization by anti-commutation relations is independent of the basis.

Compatibility. When the context is clear we will abbreviate X ∈ G(V ) as X ∈ G. Let X ∈ G(V ) and Y ∈ G(W ), we say that they are compatible if the linear map Z : V ⊕ W → A given by Z(v) = X(v) if v ∈ V and Z(w) = Y (w) if w ∈ W , extends to a homomorphism Z : Λ(V ⊕ W ) → A. In this case we write Z ∈ G(V ⊕ W ) or briefly (X, Y ) ∈ G. Note that we have a super- algebra isomorphism Λ(V ⊕W ) ≈ Λ(V )⊗Λ(W ) where tensor product is in the sense of super-algebras. From this isomorphism we get in particular Z(F )= mA[(X ⊗ Y )(F )] for all F ∈ Λ(V ⊕ W ) where mA : A⊗A→A denotes the multiplication of A. Compatibility can of course be defined for X1,...,Xn ∈ G in a corresponding way. We shall express that X1,...,Xn ∈ G are compatible by writing (X1,...,Xn) ∈ G.

8 Remark 2 The notion of compatibility which we introduce here is not standard in algebraic probability. A related concept is that of kinematic independence, see e.g. the review of Accardi [1]. Given X, Y ∈ G(V ) which are compatible we define X + Y ∈ G(V ) as

2 (X + Y )(F )= F (X + Y ) := mA[(X ⊗ Y )∆F ], F ∈ ΛV, k where mA : A⊗A→A, k ∈ N, is defined as k mA(a1 ⊗···⊗ ak)= a1 ··· ak. (2) 2 Note indeed that by compatibility Z = mA ◦ (X ⊗ Y ) : Λ(V ⊕ V ) ≈ ΛV ⊗ ΛV → A is an algebra homomorphism and therefore (X + Y )(F G)= Z(∆(F G)) = Z(∆F ∆G)= Z(∆F )Z(∆G) = (X + Y )(F )(X + Y )(G). The notation is justified from the fact that (X + Y )(v) = X(v)+ Y (v) for all v ∈ V . Similarly for any λ ∈ C we can define λX as the only homomorphism such that (λX)(v)= λX(v) for all v ∈ V .

Independence. If (X1,...,Xn) ∈ G(V1 ⊕···⊕ Vn) are compatible Grassmann variables with values in the probability space (A,ω), then we say that X1,...,Xn are (tensor) independent (with respect to the state ω) if, for all Fj ∈ ΛVj , we have that [51]

k k ω X (F ) = ω(X (F )).  j j  j j j=1 j=1 Y Y   For example if (X, Y ) ∈ G(V ⊕ W ) are (compatible and) independent Grassmann random variables with values in (A,ω), then for all v ∈ V ≈ V ⊕ 0, w ∈ W ≈ 0 ⊕ W we have vw ≈ ((v ⊕ 0) ∧ (0 ⊕ w)) ∈ Λ(V ⊕ W ) ≈ ΛV ⊗ ΛW and ω((X ⊗ Y )(vw)) = ω(X(v)Y (w)) = ω(X(v))ω(Y (w)). Let now X ∈ G(V ) be a Grassmann random variable on the probability space (A,ω). By the GNS con- struction we can construct a Hilbert space K and a vector Ω ∈ K such that there is X˜ ∈ Hom(ΛV ; L(K)) such that ω(X(F )) = hΩ, X˜(F )ΩiK. By restriction, we can always take K = X˜(ΛV ) while keeping this relation. Here the bar denotes closure with respect to the Hilbert space topology of K. Then on K we can introduce by density an involution R that acts as RX˜(F )Ω = X˜((−1)|F |F )Ω where | · | is the grading on ΛV . We conclude that, without loss of generality, as long as we are interested only in the law of a Grassmann random variable X ∈ G(V ), we can assume that it is defined on L(H) for some Hilbert space H with a vector state ω(·)=(Ω, ·Ω) and that it comes with an involution RX : H → H such that |F | RX X(F )= X((−1) F )RX for all F ∈ ΛV .

As a consequence we can always arrange to realize two Grassmann variables on the same probability space in such a way that they are compatible and independent while preserving their law. Indeed for i =1, 2, consider the Grassmann variable Xi :ΛVi → Ai defined on the probability space (ωi, Ai = L(Hi)) with involution Ri = RXi and vector state ωi = (Ωi, ·Ωi). Let H = H1 ⊗ H2, and ω : L(H) → C given by ω(x) := (Ω1 ⊗ Ω2, xΩ1 ⊗ Ω2)H, x ∈ L(H). Moreover we define the random variables X˜i : ˜ ˜ Λ(V ⊕ V ) → L(H), obtained by extending the relations X1(v)= X1(v) ⊗ R2 and X2(w)= IH1 ⊗ X2(w) for v, w ∈ V . Then X˜1(v)X˜2(w) = −X˜2(w)X˜1(v) and since Λ(V ⊕ V ) ≈ ΛV ⊗ ΛV we have also that the map X˜(v ⊕ w) = X˜1(v ⊕ 0) + X˜2(0 ⊕ w) with v, w ∈ V can be extended to a homomorphism from Λ(V ⊕ V ) to A. Hence the variables X˜i, i = 1, 2, are compatible. Moreover they have the same law as Xi because ω(X˜i(Fi)) = ωi(Xi(Fi)) for all F1, F2 ∈ ΛV , i = 1, 2. Finally they are independent: ω(X˜1(F1)X˜2(F2)) = ω1(X1(F1))ω2(X2(F2)).

2.2 Topology and calculus on Grassmann variables Before beginning we want to specify the topology we consider on G(V ). We consider here only the case where V is finite dimensional, since there is a more or less unique natural topology on G(V ). Some topologies of G(V ) when V is infinite dimensional are discussed in Section 4.

9 When V is finite dimensional all norms on V are equivalent. For this reason we choose the norm induced by the pre-Hilbert space inner product h·, ·i (that in the case of V finite dimensional is a Hilbert space inner product on V ) related to the construction of V -valued Gaussian random variables X (see Section 2.3). In this case G(V ) has a natural metric topology given by the distance

dG(V )(X, Y ) := kX − Y kG(V ) = sup kX(v) − Y (v)kA, (3) v∈V,|v|V =1 where k·kA is the natural norm in the ∗-algebra A.

Remark 3 In principle the definition of the distance (3) and the related norm k·kG(V ) does not use in any way the fact that V is finite dimensional. For this reason in the following, when V is a (in general infinite dimensional) per-Hilbert space we will use the notation kXkG(V ) for the quantity supv∈V,|v|V =1 kX(v)kA.

When V is finite dimensional G(V ) is a complete metric space with respect to dG(V ), in fact we have:

Lemma 4 The metric dG(V ) makes G(V ) a complete metric space.

Proof The quantity dG(V )(X, Y ) as defined in (3) satisfies the usual properties of a metric. If (Xn)n is a Cauchy sequence with respect to dG(V ), there is an element X ∈ L(V, A) such that (Xn)n converges to X in L(V, A). The only thing that we have to prove is that the linear map X ∈ L(V, A) can be extended to an homomorphism from ΛV to A. This is equivalent to prove that for any v, w ∈ V we have X(v)X(w) = −X(w)X(v) where the product is the natural one in A. On the other hand by the continuity of the product of A with respect to k·kA we have

X(v)X(w) = lim Xn(v)Xn(w)= − lim Xn(w)Xn(v)= −X(w)X(v). n→+∞ n→+∞ ✷

Remark 5 The distance dG(V ) is not the unique reasonable choice in G(V ) and it depends on the topology chosen on A (in this case we choose the topology of the uniform converge of operators). An essential choice to preserve the non-linear structure of G(V ), exploited in the proof of Lemma 4, seems to be that the product on A is continuous with respect to this topology.

In the following we will denote kY kG(V ) for Y ∈ G(V ) the positive number

kY kG(V ) := kY kL(V,A) = sup kY (v)kA, v∈V,kvk=1 i.e. the norm of the restriction of Y to L(V, A). This is useful to formulate a Taylor formula on G(V ). We will use the simpler notation m for the multiplication mA in A.

Consider G ∈ ΛV and define the right derivative ∂R :ΛV → ΛV ⊗ V by

∂RG := (I ⊗ ΠV )(∆G):ΛV → ΛV ⊗ V (4) where ΠV :ΛV → V is the projection from the tensor algebra ΛV onto V . Then

2 2 G(X + Y ) − G(X) − mA[(X ⊗ Y )(∂RG)] = mA[(X ⊗ Y )(I − I ⊗ ε − I ⊗ ΠV )∆G]

2 for G ∈ ΛV , where we used that G(X)= mA[(X ⊗ Y )(I ⊗ ε)∆G]. Given that

(I − I ⊗ ε − I ⊗ ΠV )∆G ∈ ΛV ⊗ Λ>2V where Λ>2V denotes the subspace of ΛV of elements of degree > 2. We can define recursively the k+1-th derivative as k+1 k ∂R = (∂R ⊗ I ⊗···⊗ I)∂R :ΛV → ΛV ⊗ V ⊗···⊗ V

10 k where ∂R is the k-th order derivative. Note that the right derivative ∂R :ΛV → ΛV ⊗ V satisfies

n n−k ∂R(f1 ··· fn)= (−1) (f1 · · ·f 6 k ··· fn) ⊗ fk, f1,...,fn ∈ V. kX=1

Of course one can define also a left derivative ∂L = (ΠV ⊗ I) ◦ ∆:ΛV → V ⊗ ΛV with similar properties and n k−1 ∂L(f1 ··· fn)= (−1) fk ⊗ (f1 · · ·f 6 k ··· fn) , f1,...,fn ∈ V. Xk=1 We will consider ΛV ⊗ V as a ΛV -bimodule and in particular we define the bilinear form h·, ·i : (ΛV ⊗ V ) ⊗ V → ΛV by hf ⊗ v, wi = fhv, wi, f ∈ ΛV,v,w ∈ V. (5) Note that there is no ambiguity on whether h·, ·i denotes this bilinear form or whether it denotes the scalar product of V .

k Remark 6 The definition of derivative ∂R and the bilinear for (5) do not depend on the fact that V is k finite dimensional. For this reason we can define ∂R and the bilinear for (5) also when V is an infinite dimensional pre-Hilbert space.

Remark 7 Note that if Godd ∈ ΛoddV then ∂RGodd = ∂LGodd. On the other hand, if Geven ∈ ΛevenV then ∂RGeven = −∂LFeven. In general, if we denote by πodd, πeven respectively the projection onto the odd and even parts of ΛV , then ∂R = (πeven − πodd)∂L. We can now state the version of Taylor formula which we will use below.

Lemma 8 Consider G ∈ ΛV and let X, Y ∈ G(V ) be two compatible Grassmann random variables such that kY kG(V ) 6 1, then

n 1 G(X + Y )= G(X)+ mk+1[(X ⊗ Y ⊗···⊗ Y )(∂k G)] + O(kY kn+1 ) (6) k! A R G(V ) kX=1 Proof We have

2 2 G(X + Y )= mA[(X ⊗ Y )(∆G)] = mA[(X ⊗ Y )(1 ⊗ ΠΛk V )(∆G)], kX>0 where the sum is finite since F ∈ ΛV is a finite polynomial. Equation (6) easily follows as soon as we prove that k+1 k 2 mA [(X ⊗ Y ⊗···⊗ Y )(∂RG)] = (k!)mA[(X ⊗ Y )(I ⊗ ΠΛk V )(∆G)]. We have

k ∂RG = (∂R ⊗ I ⊗···⊗ I) ··· (∂R ⊗ I)∂RG = (1 ⊗ ΠV )∆ ⊗ I ⊗···⊗ I ··· (I ⊗ ΠV )∆ ⊗ I I ⊗ ΠV ∆G = (I ⊗ ΠV ) ⊗ I ⊗···⊗ I (∆ ⊗I ⊗···⊗ I) ··· (I ⊗ ΠV ) ⊗ I (∆ ⊗ I) I ⊗ ΠV ∆G = (I ⊗ ΠV ) ⊗ I ⊗···⊗ I ··· (I ⊗ ΠV ) ⊗ I I⊗ ΠV (∆ ⊗ I⊗···⊗ I) ··· (∆ ⊗ I)∆G = I ⊗ ΠV ⊗ ΠV ⊗···⊗ Π V (∆ ⊗ I ⊗···⊗I ) ··· (∆⊗ I)∆G.  Let N := {1,...,n}, #N = n. For I = {j1,...,jk}⊂ N, we put vI := vj1 ··· vjk for vj ∈ V, j ∈ N. Let sgn(N\I) be a sign such that vN = sgn(N\I)vN\I vI By definition of the coproduct ∆ in the exterior algebra ΛV , we have

(I ⊗ ΠV ⊗···⊗ ΠV )(∆ ⊗ I ⊗···⊗ I) ··· (∆ ⊗ I)(∆)(v1 ··· vn)= = sgn(N\I)v ⊗ v ⊗···⊗ v . j1,...jn∈N N\{j1,...,jn} j1 jn P

11 On the other hand

(I ⊗ ΠΛk V )∆v1 ··· vn = (I ⊗ ΠΛk V ) I⊂N sgn(N\I)vN\I ⊗ vI 1 = sgn(N\I)vN\{j ,...,j } ⊗ v{j ,...,j }. k! j1,...,jk P 1 k 1 k By linearity, we therefore have P

k k (I ⊗ mΛV )(∂RG)= k!(I ⊗ ΠΛk V )(∆G).

k ⊗k ⊗k where mΛV : V ⊂ (ΛV ) → ΛV is defined as in equation (2). Since X, Y are algebra homomorphisms they commute with the product, hence we have

k+1 k 2 k k mA [(X ⊗ Y ⊗···⊗ Y )(∂RG)] = mA[(X ⊗ Y )(I ⊗ mΛV )(∂RG)] 2 = k!mA[(X ⊗ Y )(I ⊗ ΠΛk V )(∆G)].

This concludes the proof. ✷

We want to give now some precise bounds on the norms kG(X)kA and kG(X) − G(Y )kA, where G ∈ ΛV and X, Y ∈ G(V ) are two compatible Grassmann random variables. First we introduce a dim(V ) n suitable norm k·kπ on ΛV that is the norm induced by the projective norm on n=0 V . Note n n that there exists an injection iΛnV :Λ V → V given by the unique linear extension of the following relation L N N 1 σ i n (v ∧···∧ v )= (−1) v ⊗ v ⊗···⊗ v . v ,...,v ∈ V. Λ V 1 n n! σ(1) σ(2) σ(n) 1 n σX∈Sn dim(V ) dim(V ) n We then define the map iΛV := n=0 iΛnV . On n=0 V we consider the projective norm k·kπ defined as follows. If f ∈ n V then L L N N p p k k k k kfkπ := inf kf1 kV ···kfn kV , where f = f1 ⊗···⊗ fn . ( ) Xk=1 Xk=1 dim(V ) n dim(V ) For a general element g ∈ n=0 V we put kgkπ := k=0 kΠ⊗nV (g)kπ, and for any w ∈ ΛV , L N P kwkΛπV := kiΛV (w)kπ. (7)

Since iΛV is an injection, k·kπ defines a norm on ΛV . If G ∈ ΛV we define

deg(G) = max {n ∈ N, such that kΠΛnV (G)kπ 6=0} .

Let W a vector space and introduce the symmetrizer S : T W → T W as the operator from the tensor algebra generated by W in itself that is the unique linear extension of 1 S(w ⊗···⊗ w )= w ⊗···⊗ w , w ,...,w ∈ W. 1 n n! σ(1) σ(n) 1 n σX∈Sn It is important to note that for S((v + w)⊗n) the binomial formula holds, i.e. we have

n n S((v + w)⊗n)= S(v⊗k ⊗ w⊗(n−k)). (8) k Xk=0  

n Lemma 9 If Gn ∈ Λ V and X, Y ∈ G(V ) are two compatible Grassmann random variables we have that n! mk+1[(X ⊗ Y ⊗···⊗ Y )(∂k G )] = mn (S(X⊗(n−k) ⊗ Y ⊗k)(i (G ))). A R n (n − k)! A ΛV n

12 Proof For any λ ∈ R, by Lemma 8 we have

n λk G (X + λY )= G (X)+ mk+1[(X ⊗ Y ⊗···⊗ Y )(∂k G )]. (9) n n k! A R n Xk=1 On the other hand, by the binomial formula (8), we have

n ⊗n Gn(X + λY ) = mA(S((X + λY ) )iΛV (Gn)) n n n k ⊗(n−k) ⊗k = mA λ S(X ⊗ Y )iΛV (Gn) . (10) k ! Xk=0   By comparing the expressions (9) and (10) as polynomials in λ we get the thesis. ✷

Theorem 10 Let X, Y ∈ G(V ) be two compatible Grassmann random variables and let G ∈ ΛV then, for any n 6 deg(G) − 1, we have

deg(G) kG(X)kA 6 kGkΛπV (1 + kXkG(V )) (11) and G(Y ) − G(X) − n 1 mk+1[(X ⊗ (Y − X) ⊗···⊗ (Y − X))(∂k G)] 6 k=1 k! A R A (12) 6 C (1 + max(kXk , kY k ))deg(G)−n−1kGk kY − Xkn+1 n,deg(G) P G(V ) G(V ) ΛπV G (V ) where Cn,deg(G) > 0 is a suitable constant depending only on n and deg(G). In the case n = 0 we can choose C0,deg(G) = deg(G).

n Proof First of all we note that if (G1, . . . ., Gn) ∈ L(V, A) and G˜ ∈ V we have n ˜ ˜ N kmA((G1 ⊗···⊗ Gn)(G))kA 6 kGkπkG1kL(V,A) ···kGnkL(V,A). (13) Furthermore, if X ∈ G(V ) and G ∈ ΛV we have

deg(G) n ⊗n G(X)= mA(X (iΛnV (ΠΛnV (G)))). n=0 X

Using then the definition of kXkG(V ) and of kGkΛπV we get

deg(G) n n kG(X)kA 6 kXkG(V )kiΛnV (ΠΛnV (G))kπ 6 (1 + kXkG(V )) kGkπ. n=0 X

In general, writing Gh = ΠΛhV (G), by Lemma 9, we have

n 1 k+1 k Gh(Y ) − Gh(X) − k=1 k! mA [(X ⊗ (Y − X) ⊗···⊗ (Y − X))(∂RGh)] = h = mh S(Y ⊗h) − n S(X⊗(h−k) ⊗ (Y − X)⊗k) (i (G )) = A P k=0 k ΛV h  h    = mh h P S(X⊗(h−k) ⊗ (Y − X)⊗k) (i (G )) = A k=n+1 k ΛV h   h  k − n − 1   (14) = mh h Pk−n−1 S(X⊗(h−k) ⊗ (−X)⊗ℓ ⊗ (Y )k−n−1−ℓ A k=n+1 ℓ=0 k ℓ      P P⊗(n+1) h h−n−1 h−p ⊗p ⊗(Y − X) ))(iΛV (Gh))) = mA p=0 S(X ⊗ Y ) h ℓ + p h−n−1−p(−1)ℓ P (i (G )) ℓ=0 p + ℓ + n +1 ℓ ΛV h       P Using inequality (13) in relation (14) we get

n 1 k+1 k 6 Gh(Y ) − Gh(X) − k=1 k! mA [(X ⊗ (Y − X) ⊗···⊗ (Y − X))(∂RGh)] A 6 ki (G )k C (max(kXk , kY k ))h−n−1kY − Xkn+1 , ΛV h ΛPπ V n,h G(V ) G(V ) G(V )

13 where we can choose h−n−1 h−n−1−p ℓ h ℓ + p Cn,h = (−1) . p + ℓ + n +1 ℓ p=0 ℓ=0     X X

In the case n =0 we can get a better constant. Indeed we have h−1 ⊗k ⊗h−1−k kGh(Y ) − Gh(X)kA = S(X ⊗ Y ⊗ (Y − X)) kGhkΛπV

k=0 G(V ) X h−1 6 h(max( kXkG(V ), kY kG(V ))) kX − Y k. ✷ We now introduce the notion of function “depending on the space variable” v ∈ V and a suitable norm on the space of such functions. If F ∈ L(V, ΛV ) we can define the composition F (X)(v) := X(F (v)), X ∈ G(V ), v ∈ V.

as a linear (and continuous) map F (X): V → A. We define kF kΛπV as

kF kΛπV := kF kL(V,ΛπV ) = sup kF (v)kΛπ V . (15) v∈V,|v|=1

where in the r.h.s. we use the norm k·kΛπV defined in (7).

Remark 11 If F : V → ΛoddV we have that for any v1, v2 ∈ V we have

F (X)(v1)F (X)(v2)= −F (X)(v2)F (X)(v1). This means that F (X) can be extend to an homomorphism Hom(ΛV, A) (we shall denote this extension by F (X) too). Furthermore, since

F (X)(v1)X(v2)= −X(v2)F (X)(v1), we have that F (X) and X are compatible Grassmann random variables.

n n n n We can define also ∂RF : V → ΛV ⊗ ( V ) as ∂RF (v)= ∂R(F (v)), and deg(F ) in the obvious way.

Theorem 12 Suppose that F ∈ L(V, ΛV N) and let X, Y ∈ G(V ) be compatible Grassmann random vari- ables then we have that deg(F ) kF (X)kG(V ) 6 kF kΛπV (1 + kXkG(V )) n 1 k F (Y ) − F (X) − k=1 k! m[(X ⊗ (Y − X) ⊗···⊗ (Y − X))(∂RF )] G(V ) 6 6 C (1 + max(kXk , kY k ))deg(F )−n−1kF k kY − Xkn+1 n,deg(F ) G(V ) G(V ) ΛπV G(V ) P Proof The proof is a simple application of Theorem 10 to F (X)(v) and F (Y )(v) for any fixed v ∈ V and of the definition of the norm (15). ✷

2.3 Grassmann Gaussian variables Let now V be a real pre-Hilbert space with scalar product h·, ·i and with an antisymmetric bounded operator C : V → V . We recall that, by Remark 3 and Remark 6, we can extend the definitions of k k·kG(V ), ∂R and the bilinear form (5) from the finite dimensional case to the generic (infinite) pre-Hilbert space V .

Definition 13 A (V -)Grassmann (centered) Gaussian variable with correlation C is a random variable X ∈ G(V ) such that

ω(X(G)X(f)) = ω(X(h∂RG,Cfi)), G ∈ ΛV,f ∈ V. (16)

We also require that kXkG(V ) < ∞, i.e that the map X : V → A must be continuous with respect the topology induced on V by the pre-Hilbert product structure and the (norm) topology of A.

14 If we define ∂RX(G) = X(∂RG), with the understanding that X(g ⊗ w) = X(g) ⊗ w, g ∈ V and w ∈ ΛV , then we have the more familiar expression (similar to the bosonic counterpart)

ω(G(X)X(h)) = ω(h∂RG(X),Chi).

Note that the formula determines all the moments of the Gaussians. In particular (16) implies that ω(X(f1) ··· X(fn))=0 if n is odd and if n =2k is even

k σ ω(X(f1) ··· X(f2k)) = (−1) hfσ(2i−1),Cfσ(2i)iV (17) σ i=1 X Y where the sum runs over all the pairings σ of {1,..., 2k} and (−1)σ is an appropriate sign. Equation (17) is often called Wick’s rule. The right hand side of (17) can be written as a Pfaffian:

ω(X(f1) ··· X(fn)) = Pf hfi,Cfj iV , (18) 16i,j6n where the Pfaffian is defined, for an anti-symmetric n × n matrix M, to be zero if n =2k +1 and if n=2k as the polynomial in the entries of M which satisfy the relation

2 ( Pf Mij ) = det Mij . 16i,j6n 16i,j6n

The existence of a Gaussian variable, in the above sense, implies the inequality

2 2 2 2n 2 2 det hfi,Cfj iV = [ω(X(f1) ··· X(fn))] 6 kX(f1)k ···kX(fn)k 6 kXk kf1k ···kfnk , 16i,j6n well known in the mathematical physics literature relative to fermionic expansions e.g. see [114].

Remark 14 The averages of Gaussian variables depend only on the quadratic form (f,g) 7→ hf,CgiV , however analysis on the Grassmann algebra relies on the scalar product h·, ·iV also via the requirement that kX(f)kA 6 kfkV . In particular the realization of the Grassmann algebra as a family of bounded operators is not canonical. This is the reason we need to require the above continuity of the map X : V → A. In order to construct Grassmann Gaussian variables, we need “a place to hang the hat on”, this place will be a (canonical commuting relations) CAR algebra endowed with its vacuum state. To allow for arbitrary correlations C we can use a standard trick [128] which consists in doubling the generators of the CAR algebra with respect to the generators of the Grassmann algebra.

Lemma 15 For every antisymmetric and bounded C : V → V there exists a (V -)Grassmann Gaussian variable X with correlation C (on a suitable probability space (A,ω)).

Remark 16 It is important to note that Lemma (15) is peculiar of the Grassmannian setting and it has no equivalent in the commutative case in the following sense: if V is and pre-Hilbert space and S is a com bounded positive, self-adjoint operator there is no map X : V → A such that, for any v1,...,v2k ∈ V ,

com com ω(X (v1) ··· X (v2k)) = hSvi, vj i M∈{perfect matchingsX of {1,...,2k}} (i,jY)∈M for some C∗-algebra A with a positive state ω. Indeed in the commutative case the Gaussian variables are unbounded and so we must realize them in a algebra A of unbounded operators (for example we can p consider A = 16p<+∞ L ((Ω, F, P), C), for a suitable probability space (Ω, F, P) and ω(·)= E[·], which is only a ∗-algebra and not a normed algebra, thus not a Ce∗-algebra). e T Proof of Lemma 15 Let us consider ΛV itself as a real pre-Hilbert space with respect to the scalar product h·, ·iΛV on ΛV given by

hv1 ∧···∧ vn, w1 ∧···∧ wmiΛV = δnm det hvj , wki, 16j,k6n

15 for vj , wk ∈ V , j =1,...,n, k =1,...,m, where h·, ·i denotes as usual the scalar product on V . Note that n ⊗n h·, ·iΛV on Λ V is simply the restriction of the Hilbert scalar product on the tensor product V . Let H be the completion of ΛV with respect to h·, ·iΛV and denote by Ω the element in H which corresponds to 1 ∈ ΛV , often referred to as the vacuum vector. Denote by λ :ΛV → End(ΛV ) the left action of ΛV on itself given by λ(H)G = H ∧ G, where H, G ∈ ΛV . We show that this action extends to a representation of ΛV on H. Indeed λ(v), v ∈ V corresponds to a creation operator. Let us denote by λ(v)T the adjoint of the operator λ(v) with respect to the scalar product h·, ·iΛV . A standard computation shows that for v ∈ V we have n T ℓ−1 λ(v) w1 ∧···∧ wn = (−1) hv, wℓiw1 ∧ · · · ∧w 6 ℓ ∧···∧ wn. Xℓ=1 T T Namely λ (v)x can be expressed in terms of the left derivative as λ (v)x = hv, ∂Lxi (recall the definition given in equation (5)) for all x ∈ ΛV . In particular we have

{λ(v), λ(w)} = {λ(v)T, λ(w)T} =0, {λ(v)T, λ(w)} = hv, wi, v,w ∈ V, (19) where as usual {·, ·} denotes the anticommutator. As a consequence we obtain, for any x ∈ ΛV

T T T hλ(v)x, λ(v)xiΛV + hλ(v) x, λ(v) xiΛV = hx, {λ(v) , λ(v)}xiΛV = hv, vihx, xiΛV .

In particular λ(v), λ(v)T, v ∈ V , extend to bounded operators on H (still denoted by the same symbol). We now define X :ΛV → L(H) to be the algebra homomorphism given on V by

X(v) := λ(Cv)+ λ(v)T, v ∈ V.

Then X is extended in a natural way to ΛV (still denoted by the same symbol). Note that X(x), x ∈ ΛV , is indeed a bounded operator since it is the sum of products of bounded operators. Finally we define ω : L(H) → R to be the state on A = L(H) defined by Ω, that is

ω(T ) := hΩ,T ΩiΛV ,T ∈ L(H).

We claim that X is a Gaussian Grassmann random variable on (L(H),ω). We first show that it is Grassmann: {X(v),X(w)} = {λ(Cv), λ(w)T } + {λ(v)T, λ(Cw)} = hCv,wi + hv,Cwi = hv, CTwi + hv,Cwi =0, where v, w ∈ Λ, because by assumption CT = −C. Now we have also that

T ω(X(v)X(w)) = hΩ, λ(v) λ(Cw)ΩiΛV

= hλ(v)Ω, λ(Cw)ΩiΛV = hv,CwiΛV , where the first equality follows from the fact that λ(v)TΩ=0. More generally using that X(v)Ω = λ(Cv)Ω and the commutation relation (19) we have that, for all f1,...,f2k ∈ V

ω(X(f1) ··· X(f2k)) = ω(X(f1) ··· X(f2k−1)λ(Cf2k))

= −ω(X(f1) ··· λ(Cf2k)X(f2k−1)) + ω(X(f1) ··· X(f2k−2))hf2k−1,Cf2ki 2k−1 2k−1−ℓ ✘✘ = ··· = (−1) ω (X(f1) ···✘X(fℓ) ··· X(f2k−1)) hfℓ,Cf2ki Xℓ=1 Therefore ω(X(G)X(f)) = ω(X(h∂RG,Cfi)), G ∈ ΛV,f ∈ V. Note that, similarly ω(X(f)X(G)) = ω(X(hf, C∂LGi)). (20) ✷

Complex Gaussians. Later on we will need also complex Grassmann Gaussian variables, i.e. Gaussian variables taking values in ΛV with a complex pre-Hilbert space V whose Hermitian scalar product we

16 denote by (·, ·)V and we assume anti-linear in the left variable. As in the commutative setting, their definition poses no particular problems, however the interplay of the algebraic and analytic structure is here reflected on the fact that we need to fix a real structure κ on V . We will note by hh·, ·iiκ = (κ·, ·)V the associated bilinear form on V and by Aκ = κA∗κ the transposition of the operator A : V → V with respect to κ. Please refer to the Appendix A for the basic notions of real structures on complex Hilbert spaces.

Definition 17 Let V a complex pre-Hilbert space, κ a real structure over it and C : V → V a κ- antisymmetric (i.e. Cκ = −C) bounded linear operator. A (V, κ)-Grassmann (centered) Gaussian variable with correlation C is a random variable X ∈ G(V ) such that

ω(X(G)X(h)) = ω(X(hh∂RG,Chiiκ)), G ∈ ΛV,h ∈ V, where ΛV is the Grassmann algebra over C generated by V . We require that kXkG(V ) < ∞, i.e that the map X : V → A is continuous.

Lemma 18 For every κ-antisymmetric and bounded C : V → V there exists a (V, κ)-Grassmann centered Gaussian variable X ∈ G(V ) with correlation C (on a suitable probability space (A,ω)). Proof The construction of a complex Grassmann Gaussian proceeds as in Lemma 15 by considering the (complex) antisymmetric Fock space H associated to V with vacuum vector Ω ∈ H and creation operators (a(v))v∈V which are linear and satisfying the canonical anti-commutation relations (CAR) ∗ {a(w) ,a(v)} = (w, v)V for v, w ∈ V . Let X(v)= a(Cv)∗ + a(κv), v ∈ V, and consider the state ω(A) := hΩ, AΩiH for any A ∈ A = L(H). The verification that the bounded operators (X(v))v∈V forms a Grassmann algebra and that

ω(X(v)X(w)) = hhv,Cwiiκ, v,w ∈ V, as required, is left to the reader. ✷ White noise. A relevant example of a random variable taking values in an infinite dimensional Grass- mann algebra is d-dimensional (Gaussian) white noise, defined as follows.

Definition 19 A V -valued d-dimensional (Gaussian) white noise with correlation C : V → V is the (centered) Grassmann Gaussian variable Ξ ∈ G(L2(Rd) ⊗ V ) with correlation C˜ given by (Cf˜ )(x)= Cf(x) for all x ∈ Rd and f ∈ L2(Rd) ⊗ V ≈ L2(Rd; V ).

Take now a one dimensional white noise Ξ with values in V and let Bt(v) = Ξ(I[0,t] ⊗ v) for v ∈ V . For fixed t > 0, Bt extends as homomorphism to all ΛV and therefore Bt ∈ G(V ). Note also that 1/2 kBt(v) − Bs(v)kA 6 |t − s| kvkV , t,s > 0, v ∈ V, (21) so B(v) ∈ C(R+, A) (here C(R+, A) is the space of continuous maps from R+ to A). Note that B ∈

Hom(ΛV, C(R+, A)): this in particular implies that Bt1 ,...,Btn is a compatible family and that B =

(Bt)t∈R+ is a with continuous trajectories. We have that B0(v)=0,

ω(Bt(v))=0, ω(Bt(v)Bs(w)) = hv,Cwi(t ∧ s), t,s > 0, v,w ∈ V, where C is the correlation of the Grassmann Gaussian random variable Ξ. Increments of B are indepen- dent and higher order moments can be computed via Wick’s rule (17). Note in particular that

sup kBtkG(V ) < ∞, (22) 06t6T for all T > 0. Let us remark here that properties (21) and (22) are very different from the pathwise properties of the commutative (bosonic) Brownian motion where the random variable realizing Brownian motion is unbounded both with respect the probability space and in time (see Remark 16 for other observations on this topic).

17 3 Finite dimensional SDEs

We want to study simple SDEs taking values in ΛV as far as it is needed for the purpose of stochastic quantization, that is with additive white noise. We refrain to undertake here a general study of Grassmann SDEs, in particular no stochastic calculus will be needed below. It seems possible to devise such a calculus but we leave it for a future work. More precisely in this section we want to study SDEs driven by an additive Brownian motion Bt when the linear space V is finite dimensional.

Definition 20 Let F ∈ L(V, ΛV ) and assume that F (v) is odd for all v ∈ V , and let Ψ0 ∈ G(V ) 2 be a random variable compatible and independent of the Brownian motion B ∈ G(L (R+, V )), then Ψ ∈ C0([0,T ]; G(V )) is a solution in [0,T ] to the (additive) SDE driven by B with drift F and initial condition Ψ0, if,

t Ψt(v)=Ψ0(v)+ Ψs(F (v))ds + Bt(v), t ∈ [0,T ], v ∈ V, (23) Z0 where the integral with respect to the variable s is understood in Bochner’s sense (relative to the norm on G(V ) introduced in equation (3)).

Remark 21 Note that we do not require that Ψt, Ψs are compatible for t 6= s, in particular C0([0,T ]; G(V )) 6= Hom(ΛV ; C([0,T ], A)), nor we require any compatibility of Ψ and B. It will turn out that such compatibility holds in fact for the unique solution of (23), but it is not necessary to put such restriction to formulate the notion of solution.

3.1 The Grassmann Ornstein–Uhlenbeck motion In this section we introduce the Grassmann analog of the Ornstein–Uhlenbeck process, that is when F (v)= Av with A : V → V is a linear operator on V . In this case we can write down an explicit formula for the solution to equation (23) extended to t > 0.

Proposition 22 Suppose that A : V → V and F (v) = Av. The unique solution to equation (23) is given, in this case, by At A(t−·) Ψt(v)=Ψ0(e v) + Ξ(I[0,t](·)e v), t ∈ R+, (24) where Ξ is the Grassmann Gaussian noise related with the Brownian motion B (see Definition 19 and the discussion that follows it). Proof Let h(t, v) ∈ L2(R; V ) be given by

A(t−s) h(t, v)(s) := I[0,t](s)e v, s ∈ R.

At Then (24) reads Ψt(v)=Ψ0(e v) + Ξ(h(t, v)). An explicit computation using that

t t h(t, v)(s) − I[0,t](s)v = Ah(r, v)(s)dr = h(r, Av)(s)dr, s ∈ R, Z0 Z0 in L2(R; V ) gives t Ξ(h(t, v)) = Bt(v)+ Ξ(h(r, Av))dr. Z0 At At Moreover, by the definition of exponential of a matrix, ∂tΨ0(e v)=Ψ0(e Av) so

t t Ar Ψt(v)=Ψ0(v)+ Ψ0(e Av)dr + Bt(v)+ Ξ(h(r, Av))dr Z0 Z0 t =Ψ0(v)+ Bt(v)+ Ψr(Av)dr Z0 as required by equation (24) for our assumption on F . ✷

18 Remark 23 In particular this shows that if (Ψ0, Ξ) is a Grassmann Gaussian process then also the solution Ψ to the SDE (23), is a Grassmann Gaussian process compatible with (Ψ0, Ξ). In order to prove this fact it is sufficient to prove that any product of the form

ω(Ψt1 (v1) ··· Ψt2k (v2)) (25) can be computed using the products (weighted with a suitable sign due to the anti-commutation of Ψ) of the covariance ω(Ψti (vi)Ψtj (vj )) for any i, j =1,..., 2k. This property of the expectation (25) follows from the fact that

ω((Ψt1 (v1) − Ψ0) ··· (Ψt2k (vn) − Ψ0)) = ω(Ξ(h(t1, v1)) ··· Ξ(h(t2k, v2k))) = min(t ,t ) σ k i j A(ti−s) A(tj −s) = σ(−1) i=1 0 he vσ(2i−1),Ce vσ(2i)iV ds where C is the covarianceP of Ξ.Q UsingR the fact that Ψ0 is a Gaussian random variable independent of A(t−s) Ξ, and so Ψ0(e v) is also a Gaussian random variable independent of Ξ, we obtain the Gaussian behavior of Ψt. The compatibility of Ψt with (Ψ0, Ξ) follows from the fact that Ψ is a linear function of (Ψ0, Ξ). Let us now study the family of random variables

s A(t−·) Ψt (v) = Ξ(e v), t ∈ R. In the next proposition, we will show that (under some specified assumptions on A) this represents the s stationary solution to the SDE (24) (this motivates our putting the upper index s to Ψt ).

Proposition 24 Assume that all eigenvalues of A have strictly negative real part less or equal than −λA, s where λA > 0. Then, we have supt∈R kΨt kG(V ) < +∞, furthermore for any G ∈ ΛV and any t ∈ R we get s A ω(G(Ψt )) = ω(G(X )), (26) A where X ∈ G(V ) is a Gaussian random variable with covariance CA given by ∞ ATs As CA := e Ce ds. (27) Z0 where AT denotes the transpose matrix.

s Remark 25 Expression (26) in particular shows that the independence of t of the expectation ω(G(Ψt )), s expressing the stationarity of Ψt . s Proof of Proposition 24 We have easily by the definition of Ψt :

s A(t−·) 1 kΨt kG(V ) . kI(−∞,t](·)e kL2(R) . , λA s where . stands for inequality modulo some appropriate positive constant. The random variable Ψt is Gaussian so it is completely characterized (in term of moments) by its covariance. Note that the s covariance of Ψt can be easily computed as follows s s A(t−·) A(t−·) ω(Ψt (v)Ψt (w)) = ω(Ξ(I(−∞,t](·)e v)Ξ(I(−∞,t](·)e w)) = 0 A(t−·) A(t−·) −ATτ −Aτ = hI(−∞,t](·)e v, CI(−∞,t](·)e wi = he v,Ce widτ = hv, CAwi Z−∞ with CA given by (27). The appearance of the transposition is due to the fact that V is a real pre-Hilbert space. ✷ Let us point out that T A CA + CAA = −C (28) by a simple integration by parts applied to (27).

s Remark 26 Suppose that A commutes with C, then we have that Ψt is for all t ∈ R a Grassmann Gaus- T −1 sian process with correlation CA = −(A + A ) C (the operator A is invertible since all its eigenvalues has strictly negative real part).

19 3.2 Existence and uniqueness for general drift

Theorem 27 For any Ψ0 ∈ G(V ) compatible with a given Brownian motion B, there exists T > 0 such that equation (23) (for general odd F ) admits a unique solution Ψ ∈ C0([0,T ]; G(V )). Moreover Ψ is compatible with Ψ0 ⊕ B.

0 Proof We are going to construct a solution via Picard’s iteration. Let Ψt =Ψ0 for all t > 0 and define the n +1-th Picard’s iteration by

t n+1 n Ψt (vα)=Ψ0(vα)+ Ψs (F (vα))ds + Bt(vα), t ∈ [0,T ], Z0 0 0 for any basis element vα ∈ V . Then Ψ ∈ C ([0,T ]; G(V )) and since Ψ0 is compatible with B we have 1 1 that, for all t > 0, Ψt belong to the Grassmann algebra generated by Ψ0 and B in A and there Ψt (vα) 1 1 is an odd element for all α. Therefore Ψt ∈ G(V ) for all t > 0 and also Ψt is compatible with Ψ0 ⊕ B. Since this is true for all t, by approximation in the operator norm we see that Ψ1 ∈ C0([0,T ]; G(V )) and 1 n 0 that Ψ is compatible with Ψ0 ⊕ B. By induction we prove then that Ψ ∈ C ([0,T ]; G(V )) for all n > 0 and that it is compatible with Ψ0 ⊕ B. n n,α Now observe that Ψs (F (vα)) is a polynomial function of (Ψs )α and therefore by a standard con- n 0 traction argument in the Banach space A we obtain that (Ψ )n converges, as n → +∞, in C ([0,T ]; A) α α for some positive T which depends only on F , on supα kΨ0 kA and on sup06t6T,α kBt kA. If we call Ψ the limit we have that Ψ ∈ C0([0,T ]; G(V )) by the first part of the proof and we deduce easily that Ψ is compatible with Ψ0 ⊕ B. ✷

Theorem 28 The solution to equation (23) exists for all times. Proof We will have global existence as soon as we can rule out explosions, that is prove that for all α t > 0 we have supα kΨt kA < ∞ Let Θt(v)=Ψt(v) − Bt(v) and extend this map to an homeomorphism of ΛV into A. Observe that

t Θt(v)=Θ0(v)+ mA[(Θs ⊗ Bs)(∆F (vα))]ds, Z0 recall that mA : A⊗A→A denotes the multiplication in A. In particular, if we consider a (linear) basis A A of ΛV denoted by (eA)A and let Θt := Θt(eA) and Bt := Bt(eA) we have

t A A B C Θt =Θ0 + hA,B,CΘs Bs ds, 0 Z XB for real coefficients (hA,B,C)A,B,C . This is a finite-dimensional, linear system of non-autonomous ODEs A in L(H) for (Θt )A (H being the Hilbert space in which the elements of A act, see Section 2), and

t A A B C kΘt kA 6 kΘ0 kA + |h| kΘs kA kBs kAds, 0 XA XA Z XB XC with |h| := supA,B,C |hA,B,C|. Therefore by Gronwall inequality

t A A C kΘt kA 6 kΘ0 kA exp |h| kBs kAds < ∞ 0 ! XA XA Z XC for all t > 0. In particular supα kΨt(vα)kA 6 supα kΘt(vα)kA + supα kBt(vα)kA < ∞ for all t > 0 and the solution to equation (23) exists for all times. ✷

20 3.3 An Itô formula for solutions of SDEs We want to prove a kind of Itô formula for the solution to equation (23). In order to do that let QC ∈ L(ΛV ⊗ V ⊗ V, ΛV ) be given by

QC (f ⊗ v ⊗ w)= hv,Cwif, f ∈ ΛV,v,w ∈ V. (29)

For F ∈ L(V, ΛV ) and G ∈ ΛV ⊗ V we define G · F ∈ ΛV by extending linearly

(f ⊗ v) · F = fF (v), f ∈ ΛV, v ∈ V, where on both sides we have multiplication of elements in ΛV .

Theorem 29 For the global solution to equation (23) we have

t 1 ω(Ψ (G)) = ω(Ψ (G)) + ω Ψ ∂ G · F + Q (∂2 G) ds (30) t 0 s R 2 C R Z0    for all G ∈ ΛV and t > 0. Proof Fix any T > 0. Note that 1/2 kBt − BskG(V ) . |t − s| , for all 0 6 s

t deg(F ) k(Ψt − Ψs) − (Bt − Bs)kG(V ) = F (Ψs)ds . (1 + kΨk[0,T ]) |t − s|, (31) Zs G(V )

where kΨk[0,T ] := supt∈[0,T ] kΨskG(V ) which, by Theorem 28 and the continuity of Ψ with respect to the time t, is finite for every T ∈ R+. By Taylor formula (see Lemma 8, Theorem 10 and Theorem 12) applied to polynomial G of degree deg(G) we have

kG(Ψr) − G(Ψs) − (Ψs ⊗ (Ψr − Ψs))∂RG+ 1 deg(F )(deg(G)−2) 3/2 (32) + 2 (Ψs ⊗ (Ψr − Ψs) ⊗ (Ψr − Ψs))(∂R∂RG) A . (1 + kΨk[0,T ]) |r − s|

deg(F ) 1/2 since by (31) kΨt − ΨskG(V ) . (1 + kΨk[0,T ]) |t − s| . In a similar way it is possible to obtain

deg(F )(deg(G)−2) 1/2 kh∂RG(Ψu), F (Ψu)i − h∂RG(Ψs), F (Ψu)ikA . (1 + kΨk[0,T ]) |u − s| (33) On the other hand, by (31), we have

r ω((Ψs ⊗ (Ψr − Ψs))∂RG) = ω s h∂RG(Ψs), F (Ψu)idu +ω((Ψs ⊗ (Br − Bs))(∂RG)) R r  = ω s h∂RG(Ψu), F (Ψu)idu (34) +ω(h∂RG(Ψs), (Br − Bs)i) R deg(F )(deg( G)−2) 3/2 +O((1 + kΨk[0,T ]) |r − s| ) In a similar way, using inequality (31) we obtain

2 2 ω((Ψs ⊗ (Ψr − Ψs) ⊗ (Ψr − Ψs))(∂RG)) = ω((Ψs ⊗ (Br − Bs) ⊗ (Br − Bs))(∂RG)) deg(F )(deg(G)−2) 3/2 (35) +O((1 + kΨk[0,T ]) |r − s| ).

Furthermore using the fact that ∂RG(Ψs) is independent of (Br − Bs) and ((Br − Bs) ⊗ (Br − Bs)) is 2 independent of ∂RG(Ψs) (this is due to the fact that Ψs is a function of {Ψ0,Bk|k ∈ [0,s]} and, moreover, B is independent of Ψ0 and has independent increments, we have:

ω(h∂RG(Ψs), (Br − Bs)i)= hω((Br − Bs)),ω(∂RG(Ψs))i =0 where on the r.h.s. we understand that averages w.r.t. the state ω are taken componentwise, and

ω[h∂R∂RG(Ψs), (Br − Bs) ⊗ (Br − Bs)iV ⊗V ]

21 = hω[(∂R∂RG(Ψs))],ω[((Br − Bs) ⊗ (Br − Bs))]i 2 = (r − s)ω[QC (∂RG(Ψs))].

By taking a partition {ti}i∈{0,...,n}, of diameter ρn → 0 as n → +∞, of [0,t] and exploiting inequality (32) we obtain

n+1 ω(G(Ψt)) = ω(G(Ψ0)) + i=0 ω(G(Ψti )) − ω(G(Ψti−1 )) n+1 = ω(G(Ψ0)) + i=1 ω((Ψti−1 ⊗ (Ψti − Ψti−1 ))∂RG) 1 P 2 1/2 + 2 ω((Ψs ⊗ (Ψr − Ψs) ⊗ (Ψr − Ψs))(∂RG)) + O(ρn ) P t = ω(G(Ψ0)) + ω 0 ∂RG(Ψu)F (Ψu)du 1 n+1  2  1/2 + 2 i=1 (ti − tiR−1)ω[QC (∂RG(Ψti ))] + O(ρn ),

deg(F ) deg(G) where the constants in O are proportionalP to (1+kΨk[0,T ]) and do not depend on the partition. 2 Taking the limit n → +∞, and so ρn → 0, and using the fact that ω[QC(∂RG(Ψs))] is continuous in s 0 (G(Ψs) being a from R+ to A since Ψ ∈ C ([0,T ], G(V ))) we obtain the thesis. ✷

3.4 Invariant measures Consider an equation of the following kind

t Ψt(v)=Ψ0(v)+ (Ψs(Av)+Ψs(F (v)))ds + Bt(v), t > 0, v ∈ V, (36) Z0 for Ψ ∈ C0([0,T ]; Hom(V, A)), F ∈ Hom(V, ΛV ), A ∈ L(V ), B a Grassmann Brownian motion with correlation C and with initial condition Ψ0 distributed as a Grassmann Gaussian independent of B with correlation CA (compare to equation (27)). Fix U ∈ ΛV even, and consider the linear functional ω˜ : A → C given by ω˜(·) := ω(·e−2U(Ψ0)), (37) where ω is, as before, is the chosen positive state on the ∗-algebra A = L(H).

Definition 30 We say that the functional ω˜ is an invariant measure for the equation (36) if for any G ∈ ΛV and any t ∈ R+ we have ω˜(G(Ψt))=ω ˜(G(Ψ0)). (38)

Remark 31 The functional is invariant in the sense of Definition 30, only if the expectations of the form (38) are constant when expressions of the form G(Ψt) are evaluated. This means in particular that only polynomials in Ψt are considered. There is a simple condition, analogous to the Fokker–Planck equation for commutative SDEs, for checking the invariance of the functional ω˜. Here it is:

Lemma 32 If Ψ is a solution to (36) and

ω˜(LG(Ψ0))=0, G ∈ ΛV, (39) for 1 LG := ∂ G · F + Q (∂2 G), R 2 C R then ω˜ is an invariant measure in the sense of Definition 30. In other words, eq. (39) is a sufficient condition for having an invariant measure.

Proof Consider G = GA = vA where (vA)A is a (finite) linear basis of ΛV . Then we have L(GA) = B B −2U(Ψ0) B κAvB for a suitable family (κA )A,B of constants in R. Write PA(t) :=ω ˜(GA(Ψt)) = ω(GA(Ψt)e ), due to definition (37). By Itô formula (30) we have that PA is the unique solution of the system of ODEs P −2U(Ψ0) −2U(Ψ0) ∂tPA(t) = ∂tω(GA(Ψt)e )= ω((LGA)(Ψt)e )

B −2U(Ψ0) B = κAω(GB (Ψt)e )= κAPB(t), (40) XB XB

22 −2U(Ψ0) with initial condition PA(0) = ω(GA(Ψ0)e ). On the other hand condition (39) implies that

B B −2U(Ψ0) −2U(Ψ0) κAPB(0) = κAω(GB (Ψ0)e )= ω(LG(Ψ0)e )=0 XB XB which means that (PA(t))A is constantly zero by uniqueness of solution to the ODEs system (40). Since the expectations of the form (38) are linear combinations of (PA(t))A the proof is complete. ✷ The following theorem gives a sufficient condition on the SDE (36) to have the invariant measure (37).

Theorem 33 For any even U ∈ ΛevenV , the SDE

t Ψt(v)=Ψ0(v)+ (Ψs(Av)+Ψs(hC∂RU, vi)ds + Bt(v) t > 0, v ∈ V, (41) Z0 with B a Brownian motion with correlation C and Ψ0 an independent Gaussian initial condition with correlation CA (defined in equation (27)) has ω˜ as invariant measure provided

T A CA − CAA =0. (42)

Ψ0 Proof The proof is a consequence of Lemma 32. Denote ω (G) := ω(G(Ψ0)) the law of Ψ0, and observe that for H ∈ ΛV ⊗ V we have the integration by parts formula

ω((Ψ0 ⊗ Ψ0)(H)) = ω(Ψ0(QCA (∂RH))), where QC is defined in (29). Then for all G ∈ ΛV and U ∈ ΛevenV we have

−2U(Ψ0) Ψ0 −2U(Ψ0) ω((Ψ0 ⊗ Ψ0)(A(∂RG)e )) = ω (QCA (∂R(A∂RGe )))

Ψ0 −2U(Ψ0) Ψ0 −2U(Ψ0) = ω (QCA ((A ⊗ 1)∂R∂RG)e ) − 2ω (QCA ((A∂RG) ⊗ ∂RU)e ). Observe that using the equality (28) we have

QCA ((A ⊗ 1)∂R∂RG) = Tr((A ⊗ CA)∂R∂RG) 1 1 = Tr((C A + ATC )∂ ∂ G)= − Q (∂ ∂ G), 2 A A R R 2 C R R where we have used the trace Tr on V given by Tr(v ⊗ w)= hv, wi. Moreover, from (42) we have also

T QCA ((A∂RG) ⊗ ∂RU) = Tr((A∂RG) ⊗ (CA∂RU)) = Tr(∂RG ⊗ (A CA∂RU)) ATC + C A ATC − C A = Tr ∂ G ⊗ A A ∂ U + Tr ∂ G ⊗ A A ∂ U R 2 R R 2 R       1 1 = − Tr(∂ G ⊗ (C∂ U)) = − h∂ G, C∂ Ui, 2 R R 2 R R T since we assumed that A CA − CAA =0 and used (28) again. Therefore 1 ω((Ψ ⊗ Ψ )(A(∂ G)e−2U(Ψ0))) = − ωΨ0 (Q (∂2 G)e−2U(Ψ0))+ ωΨ0 (h∂ G, C∂ Uie−2U(Ψ0)). 0 0 R 2 C R R R Now observing that 1 LG(Ψ )=(Ψ ⊗ Ψ )(A∂ G)+Ψ h∂ G, C∂ Ui + Q (∂ ∂ G) 0 0 0 R 0 R R 2 C R R   we deduce that −2U(Ψ0) −2U(Ψ0) ω(LG(Ψ0)e )= ω((Ψ0 ⊗ Ψ0)(A∂RG)e )+ 1 +ωΨ0 h∂ G, C∂ Ui + Q (∂ ∂ G) e−2U(Ψ0) =0. R R 2 C R R    ✷

23 3.5 Long-time behavior for small non-linearity We will now investigate the existence of solutions which are globally bounded in time. In the commutative setting, non-linear equations can have globally bounded solutions only if the non-linearity stays uniformly small or if it shows some coercivity. Like all notions of positivity, also coercivity however does not apply well in the Grassmann setting. The only kind of coercive term we have identified is a linear drift with a negative sign. This is a very mild coercive term, but it turns out to be enough provided the non-linearity is small enough.

Consider the equation

t Ψt(v)=Ψ0(v)+ (Ψs(Av)+ λΨs(F (v)))ds + Bt(v), t > 0, v ∈ V, (43) Z0 for Ψ ∈ C([0,T ]; Hom(V, A)), with λ > 0 and F ∈ Hom(V, ΛV ) where (as in Section 3.1) A is an operator with eigenvalues having strictly negative real part less then −λA < 0. In this setting we introduce the notion of stationary solution to equation (43), extending the one we defined for the linear case in Section 3.1.

s 0 Definition 34 We say that Ψt ∈ C (R, G(V )) is a stationary solution to equation (43) (extended to all t ∈ R) of norm at most K ∈ R+ if the following two conditions hold:

s 1. supt∈R kΨt kG(V )

t s s Aτ A Ψt = λ Ψτ (F (e v))dτ + Bt t ∈ R (44) Z−∞ A A(t−·) where Bt (v) = Ξ(e (v)). Hereafter we write A L(A, C) = sup kBt kG(V ), t∈R 0 0 and we denote by KK ⊂ C (R, G(V )) the set of Ψ ∈ C (R, G(V )) such that

sup kΨtkG(V ) 6 K, t∈R for some constant K > 0.

Theorem 35 Assume K > 3L(A, C) and suppose that all the eigenvalues of A have negative real part that are less or equal than −λA, then there exists λ0(K,L,F ) depending on K, L(A, C) and F such that if |λ| 6 2λ0(K,L,F ) there exists a unique stationary solution of norm at most K to equation (43). Proof It is a simple application of the Banach fix point theorem by noting that the map

t A(t−τ) A Kλ(Ψ)t := λ Ψτ (F (e v))dτ + Bt , t ∈ R, Z−∞ is a contraction in KK . Indeed, if Ψ ∈ KK we have, by Theorem 12,

t deg(F ) −λA(t−τ) kKλ(Ψ)tkG(V ) 6 |λ|kF kΛπV (1 + K) e dτ + L(A, C), Z−∞ where kF kΛπV is defined by (15). If we request that deg(F ) kF kΛπV (1 + K) 2λ0 +2L < K, (45) λA

24 (where λ0 := λ0(K,L,F )) for |λ| 6 2λ0 we have that Kλ maps KK into itself. Furthermore

|λ| deg(F )−1 kKλ(Ψ1)t − Kλ(Ψ2)tkG(V ) 6 (deg(F ))kF kΛπV (1 + K) kΨ1,t − Ψ2,tkG(V ). λA Therefore, if we take 2λ0 deg(F )−1 kF kΛπV (deg(F ))(1 + K) < 1 (46) λA the map Kλ is a strict contraction. The fact that KK 6= ∅ will be a consequence of the Lemma (37) below. ✷

Remark 36 The function λ0 in Theorem 35 can be taken to be a decreasing function of the parameters L and K.

A We introduce an approximation for the solution to equation (44). Let X ∈ G(V ) independent of Bt X 0 X and consider the element Ψ−T,t ∈ C (R, G(V )) which is Ψ−T,t = X if t 6 −T , for some T > 0, and is a solution to the SDE

t X A(t+T ) X A(t−τ) A A A(t+T ) Ψ−T,t = X(e v)+ λ F (Ψ−T,t(e v)) + Bt − B−T (e v), Z−T when t> −T .

1 1 Lemma 37 Under the assumptions of Theorem 35, take kXkG(V ) < 6 K, then, for any |λ| 6 2 λ0(K,L,F ), X we have that Ψ−T,t ∈ KK for all t ∈ R. Proof Let τ ∈ R, by Theorem 12, then we have the bound

X −λAτ A sup kΨ−T,tkG(V ) 6 e (kXkG(V ) + kB−T kG(V ))+ t∈(−∞,τ]

|λ| X deg(F ) +kF kΛπV (1 + sup kΨ−T,tkG(V )) + L(A, C). λA t∈(−∞,τ]

On the other hand by definition of λ0 we have, for any τ > 0

1 1 |λ| e−λAτ K + K + kF k (1 + K)deg(F ) + L(A, C) < K. 3 6 ΛπV λ   A X 2K Let Rτ := supt∈(−∞,τ] kΨ−T,tkG(V ) and consider the set Z = τ ∈ R : Rτ 6 3 . The set Z is non- empty, since R < 1 K when τ 6 −T , open since τ 7→ R is an increasing continuous function of τ and τ 2 τ  closed since τ → Rτ is continuous. Then we must have Z = R and

X 2 sup kΨ−T,tkG(V ) 6 K, t∈R 3 for |λ| 6 λ0. ✷

X s Theorem 38 Under the hypotheses of Lemma 37 we have that Ψ−T,t converges to Ψt in G(V ) as T → +∞, uniformly on compact subsets of R. Proof If t > −T , by Theorem 12, we have that

X s −λA(t+T ) A deg(F ) kΨ−T,t − Ψt kG(V ) . e (kXkG(V ) + kB−T kG(V ))+ |λ|kF kΛπV (1 + K) × −T −λA(t−τ) deg(F )−1 × e dτ + |λ|(deg(F ))kF kΛπ V (1 + K) × Z−∞ t −λA(t−s) X s × e kΨ−T,τ − Ψτ kdτ. Z−T

25 By Gronwall inequality (see, e.g., Theorem 1.3.2 in [129]) we obtain

K |λ|kF k (1 + K)deg(F ) kΨX − Ψsk 6 kXk + + ΛπV × −T,t t G(V ) G(V ) 3 λ  A  t × e−λA(t+T )+αt e−ατ dτ  Z−T  . e−λA(t+T )eαt(eαT − e−αt) . e−(λA−α)(t+T )

deg(F )−1 where α = |λ|kF kΛπV (deg(F ))(1 + K) . Since λA − α> 0, by the condition (46), we have that

X s −(λA−α)T kΨ−T,t − Ψt kG(V ) . e

X s uniformly in T when t>P > −T for any such fixed P ∈ R. This proves that Ψ−T,t converges to Ψt uniformly on compact sets. ✷ An important consequence of Theorem 38 is the following one.

1 Theorem 39 Under the hypotheses and notations of Theorem 35 and Theorem 38, if kXkG(V ) < 3 K, −2U(X) |λ| 6 λ0(K,L,F ) and if ω(·e ) is an invariant measure for equation (43) (in the sense of Defini- tion 30), for any t ∈ R and for any G ∈ ΛV , we have that

ω(G(X)e−2U(X)) = ω(G(Ψs)). ω(e−2U(X)) t

Proof Using the previous notation, by Definition 30 with U(X) replacing U(Ψ0), we have that, for any t > −T , −2U(X) X −2U(X) X −2U(X) ω(G(X)e )= ω(G(Ψ−T,−T )e )= ω(G(Ψ−T,t)e ). By Theorem 38 we can take the limit at the right hand side as T → +∞, obtaining

−2U(X) s −2U(X) ω(G(X)e )= ω(G(Ψt )e ).

A s A Since (by construction) X is independent of Bt and since Ψt is a function only of B we get

−2U(X) s −2U(X) ω(G(X)e )= ω(G(Ψt ))ω(e ), from which the thesis follows. ✷

Remark 40 By Theorem 33, Theorem 39 applies to the case λF (v)= λhC∂RU, vi for λ small enough.

4 Infinite dimensional SDEs

We want to study SDEs of the form

dΨt = (AΨt + λF (Ψt))dt + dBt (47) in the case where Ψ takes values in G(V ) and V = S(Rd) ⊗ Rn, or S(Rd) ⊗ Cn, or C∞(Td) ⊗ Rn for some n ∈ N0 and with A a (possibly unbounded) linear operator on V . We discuss in detail the case V = S(Rd) ⊗ Rn, but every reasoning can be easily generalized to the other cases or more generally to any nuclear space of smooth functions to Fréchet nuclear space (see Remark 41 for the reason).

The main difference between the finite and infinite dimensional case for Grassmann SDEs is that in general there is no natural topology on G(V ). The space V = S(Rd) ⊗ Rn is a Fréchet nuclear space. Since not every element of G(V ) is continuous with respect to this topology we consider here the space G−∞(V ) ⊂ G(V ), the subset of G(V ) for which X ∈ G−∞(V ) if and only if X restricted to V ⊂ ΛV is a continuous linear map from V to A. Hereafter we always consider X ∈ G−∞(V ) both as an homomorphism from the algebra ΛV into A, and as a continuous linear map from V into A.

26 Since V is a Fréchet nuclear space whose dual is nuclear the set of continuous linear maps from V into A can be identified with the tensor product V ∗⊗Aˆ = (S′(Rd) ⊗ Rn)⊗Aˆ (where V ∗ is the topological dual of V , the Schwartz space of tempered distribution S′(Rd) ⊗ Rn and the symbol ⊗ˆ in the product V ∗⊗Aˆ means that we take the closure of V ∗ ⊗ A with respect to any natural topology on V ∗ ⊗ A being all of them equivalent since V ∗ is nuclear see, e.g., Theorem 50.1 in [168]), see, e.g., Proposition 50.5 in [168]. In other words we can look upon G−∞(V ) as the subset of elements of (S′(Rd) ⊗ Rn) ⊗ A that can be extended to a homomorphism from ΛV into A.

Remark 41 The identification of the set of continuous linear maps from V into A with V ∗⊗Aˆ does not hold if V is a generic topological vector space, for example if V is a Banach space. Indeed in the case where V is a infinite dimensional Banach space the space V ∗ ⊗ A has not unique natural norm. In ∗ particular the completion V ⊗ε A with respect the injective norm (which is the weaker of them) we have ∗ ∗ V ⊗ε A ( L(V, A) (see, e.g., [143]). For this reason we prefer to work in the case where L(V, A)= V ⊗Aˆ which is for example when V is a Fréchet nuclear space. Using a reasoning similar to the one exploited in Lemma 4, it is possible to prove that G−∞(V ) is a closed subset of V ∗ ⊗ A with respect to the strong topology (of both V ∗ = S′(Rd) ⊗ Rn and A), i.e. −∞ ∗ if (Xν )ν is a net in G (V ) converging to X in the topology of V ⊗ A, then X can be extended to an homomorphism from ΛV into A. Working on G−∞(V ) is not so easy since (S′(Rd) ⊗ Rn) ⊗ A is not even a Fréchet space. For this reason we introduce now some subsets of G−∞(V ) (which are analogous to Besov spaces) which admit a stronger metric topology.

Remark 42 In what fallow we use the notation if X ∈ G−∞(V ) we write, for any k ∈ {1,...,n}, the map Xk ∈ G−∞(S(Rn)) as follows k X (f)= X(f ⊗ ek) n where {ek}k=1,...,n is the standard orthonormal basis of R . We recall here only the definition of the Besov norm for functions taking values in a Banach space, d further references and details are given in Appendix C (see also [7, 8]). Let Ki ∈ S(R ), with i > −1, be the functions corresponding to the Littlewood–Paley blocks used in the definition of Besov spaces s s d −∞ Bp,q := Bp,q(R , R), where s ∈ R and p, q ∈ [1, +∞]. If X ∈ G (V ) we define

1 k d ∆iX(x) = (∆iX (x),..., ∆iX (x)) := (X(Ki(x − ·) ⊗ ek))k=1,...,n, x ∈ R ,i > −1

n k d where {ek}k=1,...n is the standard basis of R . The function ∆iX : R → A can be identified with an k d d element ∆iX ∈ S(R )⊗Aˆ = S(R , A) and we define

1/q n q/p qis k p kXk s Rd := 2 k∆ X (x)k dx . Bp,q,n( ,A) i A  Rd  kX=1 iX>−1 Z    s −∞ Given s ∈ R, we say that X ∈ G (V ) when X ∈ G (V ) and kXk s Rd < +∞. Thus we introduce B∞,∞( ,A) the natural metric d s (X, Y ) := kX − Y k s := kX − Y k s Rd . Hereafter we always use the G (V ) G (V ) B∞,∞,n( ,A) s d s d s s d s∗ notation Cn(R , A)= B∞,∞,n(R , A) and C = C1(R , R). Moreover C will denote the topological dual of Cs.

s Lemma 43 The set G (V ) with the metric dGs(V ) is a complete metric space. Proof The proof is analogous to that of Lemma 4. ✷ The case s> 0 can be understood more easily. In this case the element X ∈ Gs(V ) can be identified k d with a vector of k continuous function X˙ : R → A. Under this identification if f = (f1,...,fn) ∈ S(Rd) ⊗ Rn we have that n k X(f)= fk(x)X˙ (x)dx, Rd kX=1 Z

27 where the integral is taken in Bochner sense. If X ∈ Gs(V ), with s> 0, we can obtain the function X˙ in d −d y−x the following way. Let a : R → R be a smooth mollifier and define ax,ε,k(y) = ε a ε ⊗ ek, where x, y ∈ Rd and ε> 0. Then we have that  k X˙ (x) = lim X(aε,x,k)= X(δx ⊗ ek), ε→0 where the limit is taken with respect to the topology of A, and where δx is the Dirac delta with unit mass d s s∗ n d n at the point x ∈ R . More generally, if X ∈ G (V ) and g ∈C ⊗R , for any sequence (gℓ)ℓ ⊂ S(R )⊗R s∗ n converging to g with respect to the natural topology of C ⊗ R we have that (X(gℓ))ℓ converges in A to a unique element X(g) := lim X(gn). n→+∞ The map X now defined on all of Cs∗ ⊗ Rn is continuous. In other words, if X ∈ Gs(V ) then it can be identified with an element of L((Cs∗ ⊗ Rn), A). Furthermore it is simple to prove, using suitable S(Rd) ⊗ Rn approximations, that if g,h ∈Cs∗ ⊗ Rn, we have

X(g)X(h)= −X(h)X(g). (48)

In particular we have

k h h k d X˙ (x1)X˙ (x2)= −X˙ (x2)X˙ (x1), x1, x2 ∈ R ,h,k ∈{1,...,n}. The relation (48) has important consequences. For any ℓ > 1, let

ℓ s∗ n s∗ n s∗ n s∗ n Λπ(C ⊗ R ) ⊂ (C ⊗ R ) ⊗π (C ⊗ R ) ⊗π ···⊗π (C ⊗ R ) n times

ℓ s∗ n ℓ | s∗ n {z } ℓ s∗ n be the closure of Λπ(C ⊗ R ) in ⊗π(C ⊗ R ) using the natural inclusion iΛℓ(Cs∗⊗Rn) of Λ (C ⊗ R ) ℓ s∗ n in ⊗ (C ⊗ R ) (see the discussion after Lemma 8 for more details). Here ⊗π is the projective tensor product of Banach spaces whose norm is defined as follows: if f ∈⊗ℓ(Cs∗ ⊗ Rn) then

k ℓ k i i i kfk ℓ s∗ Rn := inf kf k(Cs∗⊗Rn), where f = f ⊗···⊗ f . ⊗π(C ⊗ )  j 1 ℓ  i=1 j=1 i=1 X Y X    s ℓ i s∗ Lemma 44 If X ∈ G (V ) then it can be extended to a continuous homomorphism from i=1 Λπ(C ⊗ Rn) into A for any n> 0. Here ℓ Λi (Cs∗ ⊗ Rn) is considered as equipped with the product obtained i=1 π L from the standard product of ∞ Λi (Cs∗ ⊗ Rn) via the projection onto ℓ Λi (Cs∗ ⊗ Rn). i=1L π i=1 π Proof The proof is based onL the fact that X can be identified with an elementL of L(Cs∗, A) and on the relation (48). ✷ In the case s> 0, we have for example that

˙ k1 ˙ kℓ d X((δx1 ⊗ ek1 ) ∧···∧ (δxℓ ⊗ ekℓ )) = X (x1) ··· X (xℓ), x1,...,xℓ ∈ R , k1,...,kℓ ∈{1,...,n}. We want now to define a natural notion of function F of an element X ∈ Gs(V ). For example, if s> 0, and we write Y = Fℓ(X), we want to be able to take Fn of the form

n j j ˙k kℓ kℓ Y˙ (x)= K X 1 (x + h1)X˙ (x + h2) ··· X˙ (x + hℓ), (49) (h1,...,hℓ),k1,...,kℓ k1,...,kXℓ=1 (h1,...,hXℓ)∈D where D ⊂ Rdℓ is some finite set and Kj ∈ R and j ∈ {1,...,k}. When ℓ is odd, the (h1,...,hℓ),k1,...,kℓ s d important property of the function Fn defined before is that Y = F (X) ∈Ck(R , A) and furthermore for any g,h ∈Cs∗ and k,p ∈{1,...,n} we have

Y k(g)Y p(h)= −Y p(h)Y k(g) (50) Y k(g)Xp(h)= −Xp(h)Y k(g).

28 We want to consider some more general setting than the one considered in equation (49) that maintains the important property (50).

d n ℓ s∗ n ′ d n Consider Fℓ ∈ L(S(R ) ⊗ R , Λπ(C ⊗ R )), then Fℓ can be seen as an element of (S (R ) ⊗ R ) ⊗ ℓ s∗ n s d ℓ s∗ Λπ(C ⊗ R ). Using the same technique as before we can introduce the Banach space Cn(R , Λπ(C ⊗ n ′ d n ℓ s∗ n R )) ⊂ (S (R ) ⊗ R ) ⊗ Λπ(C ⊗ R ) equipped with its natural norm. In this case we will use the notation Fℓ(X) := X ◦ Fℓ, (51) d n where (X ◦ Fℓ)(v)= X(Fℓ(v)) ∈ A for any v ∈ S(R ) ⊗ R .

s d k 2j+1 s∗ n Definition 45 We say that the linear function F ∈ Cn R , j=0 Λπ (C ⊗ R ) is a standard odd  r s d k 2j+1 s∗ n function of degree 2k +1 if there is a sequence of functions F L∈Cn R , j=0 Λ (C ⊗ R ) (where 2j+1 s∗ n s∗ n s∗ n Λ (C ⊗R ) = (C ⊗ R ) ⊗···⊗ (C ⊗ R ) and ⊗ is the algebraic tensorL product) such that for any 2j + 1 times r k 2j+1 s∗ Rn g ∈ V we have F (g|) → F (g), as{zr → +∞, in }j=0 Λπ (C ⊗ ) equipped with its standard topology. L Remark 46 When s> 0, an example of a function F = (F 1,...,F n) satisfying Definition 45 is

n k ˙ k ˙ k1 ˙ k2 F (X)(δx)= X (x) Gk1,k2 (x − y)X (y)X (y)dy, Rd k1X,k2=1 Z s+ǫ where Gk1,k2 ∈ C are functions that decrease faster than any polynomial at infinity. We note that a k finite dimensional approximation of F (X)(δx) is given by

n k,r ˙ k ˙ k1 ˙ k2 F (X)(δx)= X (x)X (x − xi)X (x − xj ) Gk1,k2 (y)dy, Di k1X,k2=1 DXi∈Pn Z d where (Pr)r is sequence of increasing partitions of R for which

max sup Gk1,k2 (y)dy → 0 k1,k2=1,...,n i ZDi as r → ∞.

s d k 2j+1 s∗ n Theorem 47 Let s ∈ R, let F ∈ Cn R , j=0 Λπ (C ⊗ R ) be a standard odd function of degree s d s d 2k +1 (see Definition 45), and consider XL∈ G (S(R )). Then F(X) ∈Cn(R , A), furthermore for any s∗ n v1, v2 ∈C ⊗ R we have

F (X)(v )F (X)(v )= −F (X)(v )F (X)(v ) 1 2 2 1 (52) F (X)(v1)X(v2)= −X(v2)F (X)(v1).

s d Proof The fact that F (X) ∈ Cn(R , A) follows from Remark 44. The relations (52) are obvious when s d k 2j+1 s∗ n r F ∈ Cn R , j=0 Λ (C ⊗ R ) . Using a finite dimensional approximation F converging to F , ✷ required by DefinitionL 45, the relations (52) are extended to the more general case.

Remark 48 An important consequence of Theorem 47 is that if X ∈ Gs(V ) and F satisfies the hypothe- ses of Theorem 47, F (X) can be extended to a Grassmann random variable (which in the following we denote by the same symbol F (X)) in Gs(V ). Furthermore, F (X) and X are compatible (in the sense of the corresponding definition in Section 2). We are now in the position to give a precise meaning to equation (47) and to the various hypotheses on its data. Let the operator A be the (generally unbounded) generator of a strongly continuous and

29 At 2 d exponentially contractive semigroup (e )t>0 on L (R ). Moreover consider the stationary Ornstein- Uhlenbeck motion A A(t−·) d n Bt (v) = Ξ(I(−∞,t] ⊗ e v), t ∈ R, v ∈ S(R ) ⊗ R , (53) ∞ d n where Ξ is a Grassmann Gaussian noise on G(C (R+) ⊗ S(R ) ⊗ R ) with covariance I ⊗ C (we consider ∞ d n ∞ on Cc (R+) ⊗ S(R ) ⊗ R the pre-Hilbert space structure given by the natural inclusion of Cc (R+) ⊗ d n 2 d n S(R ) ⊗ R in L (R+ × R ) ⊗ R ).

Definition 49 We say that the process Ψ ∈ C0([0,T ], Gs(V )) satisfies equation (47) with initial condition s Ψ0 if F is a standard odd function of finite degree, Ψ0 ∈ G (V ) is independent of Ξ, and for any t ∈ [0,T ] and any v ∈ V we have

t At A(t−s) A A At Ψt(v)=Ψ0(e v)+ λ F (Ψs)(e v)ds + Bt (v) − B0 (e v), (54) Z0 where the integral in the variable s is taken in Bochner sense with respect the norm of A.

s d k 2j+1 s∗ n Hereafter, we introduce the following notations if F ∈Cn R , j=0 Λπ (C ⊗ R ) :   k L kF kCs ,π := sup kF k s Rd k 2j+1 s∗ , n C ( ,L Λπ C ) k=1,...,n j=0

deg(F ) := 2k +1.

s d k 2j+1 s∗ n Theorem 50 Suppose that F ∈Cn R , j=0 Λπ (C ⊗ R ) and that F satisfies Definition 45, then s for any X, Y ∈ G (V ) compatible we haveL  deg(F ) s s s kF (X)kG (V ) 6 kF kCn,π(1 + kXkG (V )) (55)

s s s(V ) kF (X) − F (Y )kG (V ) 6 deg(F )kF kCn,πkX − Y kG × deg(F )−1 ×(1 + max(kXkGs(V ), kY kGs(V ))) . (56)

k,r s d k 2j+1 s∗ k 2j+1 s∗ Proof Let F ∈Cn R , j=0 Λ C (where j=0 Λ C is the usual algebraic antisymmetric tensor product) be the approximationL of F requiredL in Definition 45, then using the same reasoning of Theorem 10 and Theorem 12, if Ki is the function corresponding to the Littlewood–Paley block ∆i, we obtain

k,r k,r deg(F ) s∗ s sup kF (X)(Ki(x − ·))kA 6 sup kF (Ki(x − ·))kΛπC (1 + kXkG (V )) . (57) k=1,...,n k=1,...,n

If we multiply both sides of (57) by 2is, take the limit as r → +∞ and afterwords the supremum on x and j we get inequality (55). In a similar way we get (56) by generalizing the proof of Theorem 10 and Theorem 12. ✷

Remark 51 Hereafter we use the notations, for x > 0,

deg(F ) s fF (x)= kF kCn,π(1 + x) , (58)

deg(F )−1 s gF (x) = deg(F )kF kCn,π(1 + x) . (59) We introduce some conditions on A and Ξ in order to prove a result on existence and uniqueness of solutions to (54).

Definition 52 We say that the unbounded operator A on L2(Rd), and the Grassmann Gaussian noise Ξ with covariance C are adapted to the space Gs(V ), for a given s ∈ R, if

30 s d At −λAt i. The operator A generates a contraction semigroup on C (R ) such that ke kL(Cs(Rd),Cs(Rd)) . e , t > 0, for some strictly positive constant λA > 0. A s A ii. The Gaussian process B defined by (53) belongs to C(R, G (V )) and furthermore we have supt∈R kBt kCs(Rd,A) < +∞.

Theorem 53 Suppose that F satisfies Definition 45 and A, Ξ satisfy Definition 52 and let Ψ0 be a variable in Gs(V ) independent of the Grassmann Gaussian noise Ξ. Then there exists a T > 0 such s that there is a unique solution Ψ ∈ C([0,T ], G (V )) to equation (54). Furthermore Ψt is compatible with (Ψ0, Ξ, Ψs) for every t,s ∈ [0,T ]. Proof The proof is a straightforward generalization of that of Theorem 27 for the case of our infinite dimensional setting. ✷ In order to obtain solutions for all times it is not possible to extend directly Theorem 28, since its proof was essentially based on the hypothesis that V is finite dimensional. On the other hand the proofs of Theorem 35, Lemma 37 and Theorem 38 do not depend on the dimension. For this reason we introduce s s the space KK ⊂ C(R, G (V )) defined as the set of functions Ψt ∈ C(R, G (V )) such that

sup kΨtkCs(Rd,A) < K, t∈R for some constant K > 0, the function L given by A L(A, C) = sup kBt kCs(Rd,A), t∈R and the equation for stationary solutions corresponding to (54) as t s s A(t−τ) A Ψt = λ F (Ψτ (e ))dτ + Bt t ∈ R. (60) Z−∞ X X We also denote by Ψ−T,t the Grassmann process defined by Ψ−T,t = X if t 6 −T , T > 0, and such that X Ψ−T,t is the solution to equation (54) with initial condition X at time t = −T . Theorem 54 Suppose that F satisfies Definition 45 and A, Ξ satisfy Definition 52 and let K > 2L(A, C). Then there exists a function λ0(K,L,F ) such that for any |λ| 6 2λ0(K,L,F ) there exists a unique (stationary) solution to equation (60) in KK . Furthermore for any |λ| 6 λ0(K,L,F ) and for any random s A 1 X variable X ∈ G (V ), independent of Bt and such that kXkCs(Rd,A) < 3 K, we have that Ψ−T,t converges, s s as T → +∞, in G (V ) to Ψt uniformly on compact subsets of R. Proof Due to our appropriate infinite dimensional setting the proof is a straightforward generalization of Theorem 35, Lemma 37 and Theorem 38. ✷

Remark 55 The function λ0 defined in Theorem 54 has to satisfy the following inequalities 2λ (K,L,F )f (K) 0 F +2L < K, (61) λA

2λ0(K,L,F ) gF (K) < 1 (62) λA where fF , gF and λA are the quantities introduced in Remark 51 and Definition 52 respectively.

5 The Yukawa2 model We are now in position to study the stochastic quantization for an Euclidean fermionic QFT. We consider, as an example, the Euclidean counterpart of the Yukawa model describing the interaction of a scalar field with a spin 1/2 field in a 1+1 Minkowski space-time (“Yukawa2 model”) [127, 44, 22, 110, 166]. However we will not attempt here to remove the UV cutoff so the dimension of the space will not play any fundamental role. The Yukawa2 model contains already the main features of a large class of Euclidean fermionic theories. It will be clear from the considerations below that more general results are easily derivable, including other models involving fermions with ultraviolet cut-offs, at the price of heavier notation. For the sake of readability we choose to stick to a definite model.

31 5.1 Osterwalder–Schrader fields

We define the Euclidean Yukawa2 quantum field theory following Osterwalder and Schrader [128]. In the language of this paper, the Euclidean free Fermions introduced by Osterwalder and Schrader (compare with [128], page 282 and equation (3.3), page 284) are given by a quadruplet ψ˜ = (ψ, ψ¯) ∈ G(S(R2) ⊗ C4) of complex Grassmann Gaussian fields on the complex pre-Hilbert space V = S(R2) ⊗ C4 ⊆ L2(R2) ⊗ C4 with an Hermitian scalar product denoted (·, ·)V and satisfying 2 4 ω(ψ˜(f)ψ˜(g)) = hf,Kgic, f,g ∈ S(R ) ⊗ C , (63) with the standard real structure c given by the complex conjugation (i.e. cf = f¯) determining the 2 2 4 2 2 4 bilinear form hf,gic = (cf,g)V , and the correlation K : L (R ) ⊗ C → L (R ) ⊗ C given by the bounded c-antisymmetric operator

2 −1 0 ∇/ + mf K = (m − ∆) c , (64) f − ∇/ + m 0  f   2 where ∆ is the Laplacian on R , mf > 0 is a constant interpreted as the mass of the fermionic particle associated with the fermionic fields ψ˜, ∇/ := γ1∇1 + γ2∇2 with ∇i the partial with respect to 2 2 xi, i =1, 2, and with Euclidean Dirac matrices γ1,γ2 : C → C given by 0 1 0 −i γ = , γ = , 1 1 0 2 i 0     and such that γaγb + γbγa =2δab, a,b =1, 2. 2 −1/2 2 2 4 Note that K = (mf − ∆) u, with u an unitary operator in L (R ) ⊗ C . Hence the correlation matrix of the field ψ˜ can also be written as ˜ ˜ 2 −1/4 2 −1/4 ω(ψ(f)ψ(g)) = h(mf − ∆) f,u(mf − ∆) gic. Therefore ω is the state on A describing the Euclidean free fermion fields and we can realize ψ˜ as ˜ ˜ 2 −1/4 2 4 ψ(f) := ζ((mf − ∆) f), f ∈ S(R ) ⊗ C , where ζ˜ ∈ G(S(R2) ⊗ C4) is the complex Grassmann Gaussian field specified by the bounded covariance

2 4 ω(ζ˜(f)ζ˜(g)) = hf,ugic, f,g ∈ S(R ) ⊗ C . As a consequence, in this specific realization, we have ˜ ˜ 2 −1/4 2 −1/4 kψ(f)k = kζ((mf − ∆) f)k . k(mf − ∆) fkL2(R2)⊗C4 < ∞, (65) where the implicit constant in the inequality depends only on mf .

In order to describe the Yukawa2 model we need also to introduce an Euclidean boson field ϕ of 2 2 −1 mass mb > 0: it is the centered Gaussian field on R with covariance (mb − ∆) where mb > 0 is looked upon as a fixed parameter and ∆ is the Laplacian on R2. We can realize ϕ as the canonical 2 −κ 2 multiplication operator on the Hilbert space Hϕ = L Cloc (R ),ν where ν is the centered Gaussian −κ 2 −κ 2 −1 measure on Cloc (R ) (the local version of the C Besov space) with covariance (mb − ∆) and κ > 0  −κ 2 is any small positive quantity. We shall denote by ωϕ the state associated with ν and Cloc (R ).

In addition, large scale and small scale cutoffs are introduced as follows. Fix a smooth positive functions h ∈ C∞(R2) with compact support and let a be a regular enough (see below for the precise regularity) continuous function with compact support such that a(0) = 1. Take ε > 0 and let aε(x) = a(x/ε)ε−2. Then we define new Grassmann variables

2 ψ˜ε(x) := ψ˜(aε(·− x)), x ∈ R .

s 2 4 Then ψ˜ε ∈ G (S(R ) ⊗ C ) for some s> 0 (in general depending on the regularity of a see below).

Following the argument of Section 6 of [128] we define the Euclidean Yukawa2 model as follows.

32 n 2 Definition 56 The approximate fermionic Schwinger functions (ρε,h)n of the Yukawa2 model on R with infrared (IR) cutoff h and ultraviolet (UV) fermion cutoff ε are given by the Euclidean averages

˜ ˜ −Vε,h(ϕ) n ω ⊗ ωϕ(ψε(f1) ··· ψε(fn)e ) ρε,h(f1,...,fn)= −V (66) ω ⊗ ωϕ(e ε ) with potential Vε,h given by the unbounded (non-self-adjoint) operator on H⊗Hϕ, where H is the fermionic Hilbert space,

Vε,h(ϕ)= λ dxh(x)(ψ¯ε(x) · ψε(x))ϕ(x), (67) R2 Z λ ∈ R is a coupling constant and ω, resp. ωϕ are the fermion, resp. boson, states described above.

The potential Vε,h(ϕ) can be expressed in terms of the fields ψ˜ε as follows λ Vε,h(ϕ)= h(x)(ψ˜ε(x) · Jψε(x))ϕ(x)dx, 2 R2 Z where J(ψ, ψ¯) := (−ψ,¯ ψ). e Note that there is no difficulty in making sense of (66), due to the following observations (see 2 s also [128]). The map x ∈ R 7→ ψ˜ε(x) ∈ A = L(H) is C due to the presence of the regularization ε > 0 and the regularity of the Euclidean field ψ˜ given in (65). The massive Gaussian free field ϕ has −κ (local) regularity Cloc for arbitrarily small κ> 0, therefore, taking s>κ, by a standard Besov estimates −κ 2 for Banach valued functions (see e.g. [7, 8]), we have, for any fixed ϕ ∈Cloc (R )

˜ 2 − − kVε,h(ϕ)kA . λkψεk κ khϕk κ R2 . λkhϕkC κ(R2). C (A) B1,1 ( )

−κ It is also not difficult to see that the map Vε,h : C → A given by Vε,h : ϕ 7→ Vε,h(ϕ) is linear and continuous in ϕ, therefore measurable and that

−Vε,h(ϕ) −Vε,h(ϕ) ω ⊗ ωϕ(ψ˜(f1) ··· ψ˜(fn)e )= ω(ψ˜(f1) ··· ψ˜(fn)ωϕ(e )) e.g. by proving this first for a piecewise approximation of ϕ and then taking the limit. Moreover we have (by the properties of the )

−Vε,h(ϕ) kωϕ(e )kA 6 ωϕ(exp(C|λ|khϕkC−κ )) < ∞, for all λ ∈ R. We conclude that the averages in (66) are well defined since

−Vε,h −Vε,h(ϕ) |(ω ⊗ ωϕ)(ψ˜(f1) ··· ψ˜(fn)e )| 6 kψ˜(f1) ··· ψ˜(fn)ωϕ(e )kA < ∞.

−V (ϕ) By expanding the quantity ωϕ(e ε,h ) ∈ A in and integrating away the Gaussian field we obtain −Vε,h(ϕ) Vε,h ωϕ(e )= e , with a purely fermionic potential given by

2 Vε,h(ψ˜ε)= λ (hψ¯ε · ψε)(x)G(x − y)(hψ¯ε · ψε)(y)dxdy. (68) R2 2 Z( ) 2 −1 2 Here G = (mb − ∆) is the Green function of the operator (mb − ∆) and mb is the boson mass. This gives us the purely fermionic representation of the Schwinger functions (66) given by:

˜ ˜ Vε,h n ω(ψ(f1) ··· ψ(fn)e ) ρε,h(f1,...,fn)= . (69) ω(eVε,h )

Perturbation theory suggests that the boson field requires a mass renormalization aε which depends on the fermion cutoff ε > 0. In theory one would need to modify the bosonic Gaussian field covariance 2 −1 to include a mass renormalization and let ϕ have covariance (mb − ∆+ aε) instead of the original 2 −1 (mb − ∆) . However in this paper we will not discuss the removal of the UV cutoff and keep ε > 0 fixed. Therefore, for simplicity, we ignore this mass renormalization.

33 5.2 Stochastic quantization

The Grassmann SPDEs associated to the measure ω˜(·)= ω(·eVε,h ) is formally

˜ h 2 ˜ h 2 ˜ h ˜ dΨε,t = ((∆ − mf )Ψǫ,t + λ Fε,h,Y (Ψε,t))dt + dBε,t (70)

˜ h h ¯ h where Ψε,t = (Ψε,t, Ψε,t),

a∗2 ∗ hΨ¯ h G(y − ·)h(y)Ψh (y)Ψ¯ h (y)dy F (Ψ˜ h )= C ε ε,t ε,t ε,t , ε,h,Y ε,t f a∗2 ∗ hΨh G(y − ·)h(y)Ψh (y)Ψ¯ h (y)dy  ε ε,t R ε,t ε,t   with R  0 ∇/ + mf C = T , (71) f − ∇/ + m 0  f   −2 x 2 2 and where aε(x) = ε a ε , ε > 0, a : R → R is the regular enough function with compact support (see below for the more precise requirements) considered before, a∗2 = a ∗a , h is a smooth function (in  ε ε ε the beginning with compact support, later on we will consider h ≡ 1). The Brownian motion B˜ is described as follows. Let Ξ˜ be the white noise on the pre-Hilbert space 2 2 4 2 −1/2 2 4 L (R+) ⊗ S(R ) ⊗ C ⊆ L (R+) ⊗ B2,2 (R ) ⊗ C equipped with the scalar product

2 −1/4 2 −1/4 2 2 4 2 2 4 ((f,g)) = ((mf − ∆) f, (mf − ∆) g)L (R+×R )⊗C f,g ∈ L (R+) ⊗ S(R ) ⊗ C , (72) and having bounded covariance ˜ 2 −1/2 Cf = (mf − ∆) Cf . The (two-sided) Brownian motion B˜ on S(R2) ⊗ C4 is defined by

˜ 2 4 Bt(v) = Ξ(I[0,t] ⊗ v), t ∈ R, v ∈ S(R ) ⊗ C . where we use the convention 1 if t > 0and0 6 s 6 t I (s)= −1 if t< 0andt 6 s 6 0 . [0,t]   0 otherwise

We set B˜ε = aε∗B˜. According to Definition 49 the stationary version of equation (70) is the solution to the following integral problem

t ˜ h,s 2 ˜ h,s A(t−τ) ˜A 4 Ψε,t (v)= λ Fε,h,Y (Ψε,τ )(e v)dτ + Bt (v), t ∈ R, v ∈ C , (73) Z−∞ 2 where we take A = ∆ − mf and

˜A A(t−·) 2 4 Bt (v) = Ξ(I(−∞,t] ⊗ e v), t ∈ R, v ∈ S(R ) ⊗ C .

Remark 57 From now on we will make the following assumption. We fix δ > 0 and γ > 0, let α = α 2 α 2 γ +1/2+2δ and take a ∈ B1,∞(R ) ∩C (R ). Again we recall that ε> 0 will be fixed in all our study. The main theorem of this section concerns the construction of a stochastic quantization equation yielding (for |λ| small and ε > 0) the relation (69) for the Yukawa2 model with a smooth IR cut-off h (Theorem 58), resp. after the removal of h (Theorem 59).

Theorem 58 If |λ| is small enough and h is a smooth function with compact support, then, for any ∞ 2 4 k > 1, f1,...,fk ∈ Cc (R ) ⊗ C and t ∈ R, we have

˜ ˜ Vε,h ω(ψ(f1) ··· ψ(fk)e ) ˜ h,s ˜ h,s = ω(Ψε,t (f1) ··· Ψε,t (fk)). (74) ω(eVε,h )

34 We are also able to obtain the following result on the removal of the IR cutoff.

˜ s Theorem 59 If |λ| is small enough and if we denote by Ψε,t the solution to equation (73) with h ≡ 1, ∞ 2 4 we have that for any f1,...,fk ∈ Cc (R ) ⊗ C and for any t ∈ R

V ω ψ˜(f1) ··· ψ˜(fk)e ε,hn s s lim = ω(Ψ˜ (f1) ··· Ψ˜ (fk)). (75) Vε,h ε,t ε,t hn→1  ω (e n )  where (hn)n>1 is any sequence of smooth functions with compact support converging to 1 uniformly in Cγ+δ (see Remark 57 for the condition on γ,δ) on compact subsets of R2. The proof of these theorems will be presented in Section 5.4, resp. Section 5.5. These will be prepared by various estimates and results in the rest of the present section and in Section 5.3 that might have an interest on their own, also in view of other applications.

˜A δ γ 2 4 Lemma 60 We have Bε ∈C (R, C (R , A) ⊗ C ) and ˜A sup kBε,tkCγ+2δ(R2,A)⊗C4 < +∞. t∈R

1 Proof It is sufficient to prove the lemma for ε = 0 and γ = − 2 − 2δ, since for ε > 0 we have s s+α aε∗ : C →C by Theorem 91 in Appendix C. By definition of Grassmann Gaussian random variables, 2 if Ki ∈ S(R ) is the function corresponding to the Littlewood–Paley block ∆i (see Appendix C for the definition), we have

2 ˜ ˜ (∆−mf )(t−τ) 2 1/4 sup k∆iB0,t − ∆iB0,skA⊗C4 . kI(−∞,t](τ)e (−∆+ m ) (Ki)+ x∈R2

2 (∆−mf )(s−τ) 2 1/4 −I(−∞,s](τ)e (mf − ∆) (Ki)kL2(R2×R) .

2 (∆+mf )(t−τ) 2 1/4 . kI(s,t](τ)e (mf − ∆) (Ki)kL2(R2×R)+

2 2 (∆−mf )(t−s) −(∆−mf )(s−τ) 2 1/4 +kI(−∞,s](τ)(e − 1)e (mf − ∆) (Ki)kL2(R2×R) .

t 1/2 2 −2(|p|2+m2 )(t−τ) . φi(p) e f |p|dpdτ + R2 Z Zs  1/2 s t−s 2 2 −(|p|2+m2 )k 2 −2(|p|2+m2 )(s−τ) + |p| e f dk φi(p) e f |p|dpdτ . R2 Z Z−∞ Z0  ! 1/2 t−s 1 −1+4δ . 1−2δ dτ |p| dp + i−1 i τ Z2 .|p|.2 Z0  ! 1/2 t−s 2 1 1 −1+4δ i( 2 +2δ) δ + 1−δ dk |p| dp . 2 |t − s| . i−1 i k Z2 .|p|.2 Z0  ! From this the thesis follows from the definition of the norm of the Besov space Cγ(R2, A). ✷

s d Lemma 61 For any s > 0, the function Fε,h,Y is well defined on C (R , A) into itself, it satisfies 2 Definition 45. Furthermore we have kFε,h,Y kCs,π . khkCs . s 2 3 s∗ 2 4 Proof First we prove that F ∈C (R , Λπ(C (R ) ⊗ C )). We start by noticing that, if we denote by δx 2 0∗ 2 the Dirac delta with unit mass at the point x ∈ R , we have δx ⊗ ei ∈C (R ), for i =1,..., 4, where ei is the projection on the component i-th of C4. Furthermore, since for s 6∈ N, s> 0, Cs(R2) is the space of s continuous Hölder functions, we have that for any 0 < s′ < s and s′ 6∈ N the map x 7−→ δ ⊗ e is ′ ′ x i s 2 s∗ 2 4 s 2 s∗ 2 4 in C (R , C (R ) ⊗ C ). This and the fact that, by Theorem 92, C (R , Λπ(C (R ) ⊗ C )) is a Banach algebra with respect to the multiplication ∧, imply that the map

F˜(x) := h(x)(δx ⊗ e1) ∧ (δx ⊗ e3) + (δx ⊗ e2) ∧ (δx ⊗ e4)

35 ′ 2 2 s∗ 2 4 is s -Hölder as a function from R into Λπ(C (R ) ⊗ C ), with

˜ ′ kF k s R2 2 s∗ R2 C4 . khkCs (76) C ( ,Λπ(C ( )⊗ )) where the constants in the symbol . do not depend on h. On the other hand, by Theorem 90, G∗ F˜ ∈ ′ s +2 2 s∗ 2 4 ′ s 2 s∗ 2 4 C (R , Λπ(C (R ) ⊗ C )) which, if s − s > 2, is contained in C (R , Λπ(C (R ) ⊗ C )). Since every ∗2 ˜ component of Fε,h,Y is of the form ∇/ + mf aε ∗ ((h(x)δx ⊗ ei) ∧ G∗F ), using the fact that, by Theorem 90 and Theorem 91, ∇/ + m a∗2∗· : Cs(R2, A) → Cs−1+2α(R2, A) ⊂ Cs(R2, A) (since α > 1 f ε   2 for the conditions in Remark 57) is a continuous operator, and by Theorem 92 we get the thesis. As  2 byproduct of the previous reasoning we obtain also that kFε,h,Y kCs,π . khkCs . This can be seen using the fact that, when s> 0, Ψ˜ ε,t is Hölder continuous in space and that G decreases exponentially at infinity. In order to finish the proof we have to show that Fε,h,Y admits an approximation of the form required by Definition 45. This can be obtained by approximating the convolution with G and aε (the regularizing function in the definition of Ψ˜ ε) using a finite sum. Indeed, if Pn is a sequence of increasing (finite) partition of the R2, we can approximate the integral G∗F˜(x) by the sum

2 Sn(x)= F˜(x − xi) G(y)dy, x ∈ R , Di DXi∈Pn Z where xi ∈ Di is any point in Di. If we choose the sequence of partitions Pn in such a way that ′ s 2 s∗ 2 4 limn→+∞ sup G(y)dy = 0, we have that, as n → ∞, Sn → G∗F˜ in C (R , Λπ(C (R ) ⊗ C )), i Di ′ ∗2 where s

Theorem 62 Equation (70) admits a unique (local in time) solution, starting from 0 in Gγ (V ) (γ being defined before Theorem 58 and Theorem 59). Furthermore if |λ| is small enough, there exists a unique solution to the stationary equation (73) (associated with (70)) which is uniformly bounded in Cγ (Rd, A)⊗ 4 C by some constant K > 0 (that can be chosen independent of h when khkCγ < C for some C > 0). Proof The proof is based on Theorem 53 and Theorem 54, whose hypotheses are satisfied in this setting because of Lemma 60 and Lemma 61. The uniformity on the constant K and the bounds on the admissible λ follows from Lemma 61, Theorem 50 and the inequalities (61) and (62). ✷

5.3 Finite dimensional approximations in finite volume The next step in order to prove Theorem 58 and Theorem 59 is to introduce finite dimensional approxi- 2 2 2 mations of the Yukawa model. Consider the torus TR = R \(2πRZ) for R> 0 and the periodic version ψ˜R of the Fermion field ψ˜ defined by

ψ˜R(f) := ψ˜(f(·− 2πRn)), Z2 nX∈ 2 4 ˜ ˜ 2 ˜ for all f ∈ S(R )⊗C . Then letting ψR,ε(x) := ψR(aε(·−x)) we have that the function x ∈ TR 7→ ψR,ε(x) γ+2δ 2 is well defined and C (T , A). Moreover ψ˜R,ε(x) is uniformly bounded in A and ˜ ˜ kψR,ε(x) − ψε(x)kCγ+2δ (K,A⊗C4) → 0, uniformly on compacts K as R → ∞. As a consequence we have

ω(ψ˜(f ) ··· ψ˜(f )eVε,h ) ω(ψ˜ (f ) ··· ψ˜ (f )eVR,ε,h ) 1 n = lim R 1 R n (77) ω(eVε,h ) R→∞ ω(eVR,ε,h ) with 2 VR,ε,h := λ dx(hψ¯R,ε · ψR,ε)(x)G(x − y)(hψ¯R,ε · ψR,ε)(y)dxdy, R4 Z

36 2 for all compactly supported f1,...,fn,h ∈ S(R ).

2 2 2 2 Let now (ek)k∈Z2 be the orthonormal Fourier basis of (the complex) L (TR) and PN : L (TR) → 2 2 2 L (TR) the orthogonal projection to the closed subspace generated by {ek : k ∈ Z , |k| 6 N}. Let ˜ ˜ ˜ ˜ ψN,R,ε = PN ψR,ε and observe again that kψN,R,ε(x) − ψR,ε(x)kCα(R2) → 0 as N → ∞, for any α > 0. Defining 2 VN,R,ε,h := λ (hψ¯N,R,ε · ψN,R,ε)(x)Gε(x − y)(hψ¯N,R,ε · ψN,R,ε)(y)dxdy R4 Z and we easily conclude that

ω(ψ˜(f ) ··· ψ˜(f )eVε,h ) ω(ψ˜ (f ) ··· ψ˜ (f )eVN,R,ε,h ) 1 n = lim lim N,R 1 N,R n . (78) ω(eVε,h ) R→∞ N→∞ ω(eVN,R,ε,h )

2 Now consider the finite set ΛN = {k ∈ Z : |k| 6 N} and let

θ˜(k) := ek(x)ψ˜N,R(x)dx, k ∈ ΛN . T2 Z R ˜ We can show easily that (θ(k))k∈ΛN is a Grassmann Gaussian vector with correlation given by

0 (i 6k +m )−1 ω(θ˜ (k)θ˜ (ℓ)) = δ f (79) α β k+ℓ=0 − (i 6k +m )−1T 0  f α,β 4 where α, β =1,..., 4 denote the canonical coordinates in C and k,ℓ ∈ ΛN . Moreover

˜ ˜ 2 ψN,R(x)= ek(x)θ(k), x ∈ TR kX∈ΛN 2 and ψ˜N,R,ε = aε ∗T2 ψ˜N,R where the convolution is done on T as indicated. R R

At this point we have reduced via two subsequent approximations the number of degrees of freedom 4 to a finite amount, described by the Grassmann random variable θ˜ ∈ G(ΛN ⊗ C ).

We introduce the following equation

t h ,s h ,s 2 ˜ R 2 ˜ R (∆−mf )(t−τ) ˜A Ψε,N,R,t(v)= λ PN (Fε,hR,Y (PN (Ψε,N,R,τ )))(e v)dτ + Bε,R,t(v) (80) Z−∞ ˜A 2 ˜A where v ∈ V , (Bε,R,t)t∈R is the Ornstein–Uhlenbeck motion on TR obtained from (Bε,t)t∈R via the restriction from R2 to T2 of the periodicization

˜A ˜A 2 Bε,R,t(f) := Bε,t(f(·− 2πRn)), f ∈ S(R ), Z2 nX∈ and 2 hR(z) := h(z − 2πRn), z ∈ R . Z2 nX∈ Obviously we have that ˜ hR,s ˜A (I − PN )Ψε,N,R,t = (I − PN )Bε,R,t.

˜ hR,s This implies that PN (Ψε,N,R,t) solves a finite dimensional Grassmann SDE. Furthermore the drifts ˜ hR,s PN (Fε,hR,Y (PN (Ψε,N,R,t))) can be looked upon as the differentials of the function

U(PN (Ψ))˜ = hL(x)hL(y)(PN (Ψ)PN (Ψ))(¯ x)G(x − y)(PN (Ψ)PN (Ψ))(¯ y)dxdy T4 Z L ∗2 ˜ times the matrix PN aε ∗Cf (that is the covariance of the finite dimensional Brownian motion PN (Bε,R,t)). This implies the following key lemma.

37 Lemma 63 For |λ| small enough (that can be chosen in a way independent of the spatial cut-off h of Def. 56), N ∈ N, R> 0, ε> 0, t ∈ R, we have that for any k > 1 and f1,...,fk some linear combination 2 2 of the first N vectors e1,...,eN in the ONB (ek)k∈N of L (TR),

˜ ˜ VN,R,ε,h ω(ψN,R(f1) ··· ψN,R(fk)e ) hR,s hR,s = ω(Ψ˜ (f1) ··· Ψ˜ (fk)). ω(eVN,R,ε,h ) ε,N,R,t ε,N,R,t

˜ hL,s Proof The proof is based on the fact that for any linear combination fi of e1,...,eN we have Ψε,N,R,+∞(fi)= ˜ hR,s ˜ ˜ PN (Ψε,N,R,+∞)(fi). Thus the thesis follows by letting U(ψ) = VN,R,ε,h(ψ)/2 in Theorem 39 and using Remark 40. ✷

5.4 The infinite dimensional equation in finite volume Consider the equation

t h ,s h ,s 2 ˜ R 2 ˜ R (∆−mf )τ ˜A Ψε,R,t(v)= λ Fε,hR,Y (Ψε,R,τ )(e v)dτ + Bε,R,t. (81) Z−∞

This is a modification of (80) where the finite dimensional projections PN are removed from the non- linearity Fε,hR,Y .

Remark 64 Lemma 61 holds for the drift of equation (81), and it is easy to generalize Lemma 60 to ˜A the OU process (Bε,R,t)t∈R. This implies that it is possible to generalize Theorem 62 to equation (81) proving existence and uniqueness of its global solution (when |λ| is small enough). Finally, it is important ˜ hR,s to note that the bound K on the solution Ψε,R,t to equation (81), and the bounds on the constant |λ| can be chosen independent of h when khkCγ < C for some constant C > 0. Equation (81) is infinite dimensional and in order to generalize Lemma 63 we cannot apply directly the finite dimensional results contained in Theorem 39 and Remark 40. For this reason we need to prove that the solution to equation (80) converges, as N → +∞, to the solution to equation (81).

Theorem 65 For |λ| small enough, any 0 0, we have that

˜ hR,s ˜ hR,s Ψε,N,R,t → Ψε,R,t, in Gq(C∞(T2) ⊗ C4) when N → +∞, uniformly in t ∈ R.

hR,s hR,s Proof We have that the difference Ψε,N,R,t − Ψε,R,t satisfies the following inequality

hR,s hR,s t kΨ˜ − Ψ˜ kCq ε,N,R,t ε,R,t −λA(t−τ) hR,s . e k(I − P )F (P (Ψ˜ ))k q dτ + λ2 N ε,hR,Y N ε,N,R,τ C Z−∞ t −λA(t−τ) ˜ hR,s ˜ hR,s + e gF (max(kΨε,N,R,τ kCq , kPN Ψε,N,R,τ kCq )) × Z−∞ ˜ hR,s ˜ hR,s ×(kΨε,N,R,τ − PN Ψε,N,R,τ kCq )dτ + t −λA(t−τ) ˜ hR,s ˜ hR,s + e gF (max(kΨε,N,R,τ kCq , kΨε,R,τ kCq )) × Z−∞ ˜ hR,s ˜ hR,s ×kΨε,N,R,τ − Ψε,R,τ kCq dτ. Using Gronwall inequality (see, e.g., Theorem 1.3.3 in [129]) we obtain that

t ˜ hR,s ˜ hR,s 2 kΨε,N,R,t − Ψε,R,tkCq . (kI − PN kL(Cγ ,Cq )) 1+ exp(−(λA − λ gF (K))(t − τ))dτ .  Z−∞  2 Since λ is small we have that λA − λ gF (K) > 0 (this is exactly the request (62) for |λ| < λ0(K,L,F )). This means that ˜ hR,s ˜ hR,s kΨε,N,R,t − Ψε,R,tkCq . (kI − PN kL(Cγ ,Cq )).

38 On the other hand, since, γ > q by assumption, we have kI − PN kL(Cγ ,Cq ) → 0 as N → +∞. The thesis ˜ hR,s ˜ hR,s ✷ is proved noting that the convergence Ψε,N,R,t → Ψε,R,t is uniform in t. An easy corollary of Theorem 65, Lemma 63, and the convergence (78) is the following:

Corollary 66 For |λ| small enough (that can be chosen independent of h when khkCγ < C where C > 0), ∞ d 4 for any k > 1, f1,...,fk ∈ C (TL) ⊗ C , R> 0, and any t ∈ R we have

˜ ˜ VR,ε,h ω(ψR(f1) ··· ψR(fk)e ) hR,s hR,s = ω(Ψ˜ (f1) ··· Ψ˜ (fk)). ω(eVR,ε,h ) ε,R,t ε,R,t

5.5 The infinite volume limit In order to tackle the infinite volume limit and prove Theorem 58 and Theorem 59 we have to introduce d weighted Besov spaces (see the Appendix C and [169] for the details). Let ℓ ∈ R and consider ρℓ : R → R+ 2 (−ℓ/2) ′ d defined by ρℓ(x) =(1+|x| ) . Consider the following norm on S (R )⊗A (where as before A = L(H) and H is the fermionic Hilbert space)

js kuk s Rd = sup sup 2 k∆j u(x)ρℓ(x)kA. Cℓ ( ,A) j>−1 x∈Rd

s d ′ d Define the space C (R , A) as the subspace of S (R ) ⊗ A where k·k s Rd is finite. We define also ℓ Cℓ ( ,A) s s d −∞ d n d n Gℓ (V ) = Cℓ (R , A) ∩ G (V ) (recall that V = S(R ) ⊗ R or S(R ) ⊗ C for some n > 1) and the distance d s (X − Y ) := kX − Y k s Rd . Gℓ (V ) Cℓ ( ,A) Remark 67 It is important to note that, by Theorem 89 below,

sup k∆j u(x)ρℓ(x)kA ∼ sup k∆j (ρℓu)(x)kA, x∈Rd x∈Rd where ∼ means that both quantities considered can be bounded from above and from below by some positive constant times the other quantity.

Lemma 68 Let γ,δ be as stated in Remark 57, for any s<γ +2δ and any ℓ> 0 we have that

A A sup d s (B˜ , B˜ ) → 0 Gℓ (V ) ε,t ε,R,t t∈R when R → +∞. Proof Using a method similar to the one in the proof of Lemma 60, it is simple to see that ˜A sup(sup k(∆ + 1)Bε,R,tkCγ+2δ−2(R2,A)⊗C4 ) < +∞, R>1 t∈R which implies that, for any s<γ + δ, we have ˜A ˜A −i(γ+2δ) k∆i(Bε,R,t − Bε,t)k∞ 6 C12 (82) ′ where the constant C1 ∈ R+ does not depend on i and R> 1. This means that, fixing s−1 and, thus by the uniform estimate (82) and Lebesgue’s dominated convergence theorem, if, for any i > −1, A A A A k∆ (B˜ − B˜ )ρ k → 0 as R → +∞, then kB˜ − B˜ k s R2 C4 → 0 as R → +∞, and thus, by i ε,R,t ε,t ℓ ∞ ε,R,t ε,t Cℓ ( ,A)⊗ A A inequality (83), kB˜ − B˜ k s R2 C4 → 0 as R → +∞. ε,R,t ε,t Cℓ ( ,A)⊗ Fix i > 1, then we have ˜A ˜A k∆i(Bε,R,t − Bε,t)(x)ρℓ(x)kA⊗C4 6 ˜ ˜ 6 kρℓ(x)ΞR(Ki(x − ·)Gt(·)I[−κR,κR]2 ) − ρℓ(x)Ξ(Ki(x − ·)Gt(·)I[−κR,κR]2 )kA⊗C4 + ˜ ˜ +kρℓ(x)ΞR(Ki(x − ·)Gt(·)I([−κR,κR]2)c ) − ρℓ(x)Ξ(Ki(x − ·)Gt(·)I([−κR,κR]2)c )kA⊗C4 6 (84) ˜ 6 kρℓ(x)ΞR(Ki(x − ·)Gt(·)I([−κR,κR]2)c )kA⊗C4 + ˜ +kρℓ(x)Ξ(Ki(x − ·)Gt(·)I([−κR,κR]2)c )kA⊗C4 ,

39 where K˜i is given by 1/4 K˜i(x) := (∆ + mf ) (Ki),

Ki being the function related to the block i-th Littlewood-Paley, i.e. ∆if = Ki∗f, Gt is the Green function (∆+m2)t ˜ ˜ of the operator e , and where we use that ΞR(Ki(x−·)Gt(·)I[−κR,κR]2 ) = Ξ(Ki(x−·)Gt(·)I[−κR,κR]2 ) 2 1 ˜ since they coincide on [−κR,κR] for any 0 6 κ < 2 . We focus on the term J(x) := kρℓ(x)Ξ(Ki(x − ·)Gt(·)I([−κR,κR]2)c (·))kA⊗C4 . In this case we can write for |x| >κ1R +∞ 1 ˜ ˜ J(x) . ℓ sup Ki(x − y1)Ki(x − y2) × (κ R) R2 R4 1 x∈ Z0 Z

Gt(y1 − z)Gt(y2 − z)dz dy1dy2dt ([−κR,κR]2)c ! Z

1 α/2 α/2 . ℓ exp(−βi|y1| − βi|y2| )× 4 (κ1R) R+×R Z

× Gt(y1 − z)Gt(y2 − z)dz dy1dy2dt R2 Z  +∞ 2 α/2 α/2 1 (−βi|y1| −βi|y2| ) |y1 − y2| 2 dt . ℓ e exp − − 2m t dy1dy2 (κ R) R4 2t t 1 Z Z0   1 α/2 α/2 . (−βi|y1| −βi|y2| ) ℓ e ( | log(|y1 − y2|)| + 1)dy1dy2 (κ R) R4 1 Z 1 . ℓ , (κ1R)

α −βi|y| where we used the fact that there are βi (depending on i > −1) and α ∈ (0, 1) such that |K˜i(y)| . e (see Remark 88 below). Consider now |x| 6 κ1R, then we have

+∞ J(x) . K˜i(x − y1)K˜i(x − y2)P (y1,y2,t)dy1dy2dt + 0 |y1|,|y2|>κ2R Z Z

+∞ K˜i(x − y1)K˜i(x − y2)P (y1,y2,t)dy1dy2dt + 0 |y1|,|y2|<κ2R Z Z

+∞ 2 K˜i(x − y1)K˜i(x − y2)P (y1,y2,t)dy1d y2dt 0 |y1|>R,|y2|<κ2R Z Z

. J1 (x)+ J2(x)+ J3(x),

1 where κ1 <κ2 <κ< 2 and

P (y1,y2,t)= Gt(y1 − z)Gt(y2 − z)dz 2 c Z([−κR,κR] ) ! Then we get +∞ α α J1(x) . exp(−βi|x − y1| − βi|x − y1| )P (y1,y2,t)dy1dy2dt Z0 Z|y1|,|y2|>κ2R +∞ α α . exp(−βi(|y1| − |x|) − βi(|y2| − |x|) ) × Z0 Z|y1|,|y2|>κ2R

Gt(y1 − z)Gt(y2 − z)dz dy1dy2dt R2 Z  +∞ 2 α α ′ α ′ α −βi(κ2−κ1) R 2 (−βi|y | −βi|y | ) |y1 − y2| 2 dt . (e (1 + R )) e 1 2 exp − − 2m t dy1dy2 R2 2t t Z Z0   α α ′ α ′ α −βi(κ2−κ1) R 2 (−βi|y1| −βi|y2| ) . (e (1 + R )) e ( | log(|y1 − y2|)| + 1)dy1dy2 R2 α α Z . (e−βi(κ2−κ1) R (1 + R2))

40 For J2(x) we have

+∞ 2 2 J2(x) . (κ2R ) sup Gt(y1 − z)Gt(y2 − z)dz dt 2 c |y1|<κ2R,|y1|<κ2R 0 ([−κR,κR] ) ! Z Z +∞ − 2 2 2 2 (κ κ2) R − 2m2t dt . (κ2R ) e 2t t Z0 2n +∞ 2 2 (κ − κ ) 2 (κ − κ ) κ . 2 (κ R2)2 tn−1e−2m tdt . 2 2 . R2n 2 R2n−4 Z0 −n where we used that exp(−x) . x for any n ∈ N. The term J3 can be estimated in the same way combining the techniques used for J1 and J2, In conclusion

2 2 α α (κ − κ2) κ2 1 −βi(κ2−κ1) R 2 sup J(x) . 2n−4 + ℓ + e (1 + R ) → 0 R2 R (κ R) x∈  1  when n> 2 and ℓ> 0. ˜ For kΞR(Ki(x − ·)Gt(·)I([−κR,κR]2)c )kA⊗C4 we have ˜ kΞR(Ki(x − ·)Gt(·)I([−κR,κR]2)c )kA⊗C4 . +∞ ˜ ˜ (85) . 0 R4 Ki(x − y1)Ki(x − y2)PR(y1,y2,t)dy1dy2dt ,

R R where

PR(y1,y2,t) := Gt(y1 − z)Gt(y2 − z)dz + Gt(y1 − z)Gt(y2 − z)dz. T2 2 c T2 \([−κR,κR] ) ′ Z2 Z R z X∈R Z R

By noting that for |y1|, |y2| >κ2R

+∞ PR(y1,y2,t)dt 6 C2(| log(|y1 − y2|)| + 1) Z0 for some constant C2 ∈ R+ independent of R> 1 and, if one of the inequalities |y1| <κ2R or |y2| <κ2R holds, then 2 2 ′ (κ−κ2) R 2 exp −β t +2m t P (y ,y ,t) 6 . R 1 2  t  ˜ Then using inequality (85), it is possible to prove that supx∈R2 kρℓ(x)ΞR(Ki(x−·)Gt(·)I([−κR,κR]2)c )kA⊗C4 → ˜ 0 as R → +∞ by exploiting an argument analogous to the one mentioned obove for proving supx∈R2 kρℓ(x)Ξ(Ki(x− ✷ ·)Gt(·)I([−κR,κR]2)c )kA⊗C4 → 0. We now prove the analogue of Theorem 65 for the case where the space cut-off R is removed (infinite volume case):

˜ hR,s ˜ h,s q 2 4 Theorem 69 Let |λ| be sufficiently small and 0

kFε,h ,Y (Ψ˜ 1) − Fε,h,Y (Ψ˜ 2)k q R2 C4 . gF (max(khRΨ˜ 1kCq , khΨ˜ 1kCq , khΨ˜ 2kCq ))× R Cℓ ( ,A⊗ )

× (khkCq kΨ˜ 1 − Ψ˜ 2k q R2 C4 + kΨ1k q R2 C4 khR − hk q ), (86) Cℓ ( ,A⊗ ) C0 ( ,A⊗ ) Cℓ where gF is the function in Remark 51 and h the IR cut-off in Definition 56. Using a reasoning similar to the one exploited in the proof of Theorem 65 we get, for any t ∈ R,

hR,s h,s ˜ ˜ q q ˜A ˜A q kΨε,R,t − Ψε,t kC (R2,A⊗C4) . (khR − hkC + sup kBε,τ − Bε,R,τ kC (R2,A⊗C4))× ℓ ℓ τ∈R ℓ

41 t 2 × 1+ exp(−(λA − λ gF (K))(t − τ))dτ .  Z−∞  By |λ| small enough, the bound converges to 0 as R → +∞ uniformly in t ∈ R, since, by Lemma 68,

˜A ˜A lim sup kBε,τ − Bε,R,τ kCq(R2,A⊗C4) → 0 R→+∞ τ∈R ℓ and khR − hk q → 0, as R → +∞, by the definition of hR. ✷ Cℓ ˜ s Hereafter we denote by Ψε,t the solution to equation (70) with h ≡ 1 (so with the spatial cut-off h removed).

2 Theorem 70 Using the previous notations and hypotheses, let (hn : R → R)n>1 be a sequence of smooth functions with compact supports converging to 1 uniformly on compacts in the norm of Cγ and with a uniform bound in Cγ. For any 0

˜ hn,s ˜ s Ψε,t → Ψε,t

q 2 4 in Gℓ (S(R ) ⊗ C ), for any ℓ> 0 and uniformly in t ∈ R when n → +∞.

Proof The proof is analogous to the one of Theorem 69 where in inequality (86) we have khn − 1k q Cℓ (instead of khR − hk q ) which converges to 0, as n → +∞, by hypothesis, when ℓ> 0. ✷ Cℓ Proof of Theorem 58 and Theorem 59 The proof is analogous to the one of Corollary 66 where Theorem 65 is replaced by Theorem 69 and Theorem 70 respectively. ✷ These theorems provide an alternative construction of the finite, resp. infinite, volume Euclidean Yukawa model with UV cut-off as an invariant state given by the solution of an SPDE for a Grassmann valued process. In this sense also a major step in the program of the stochastic quantization for the Yukawa model has been achieved.

A Remarks on the Grassmann Gaussians

In this appendix we provide some standard facts on real structures on complex (pre-) Hilbert spaces and then discuss an alternative realization of Grassmann Gaussian variables on a tracial state.

There are various ways to constructs Grassmann Gaussians and while in the main body of the paper we used the approach of Osterwalder and Schrader, based on the CAR algebra of creation and annihilation operators on Fock space and the state generated by the Fock vacuum, here we would like to discuss other approaches based on self-dual Clifford algebras and their associated tracial state. This goes back to the work of Gross [70] and Palmer [130] and to the “real-wave” representation of the Clifford algebra introduced by Segal [154, 17] and nicely described by Gross in [68]. For more background information, aside from the already cited works [70] and [130], we refer to Araki [16, 13, 14].

A.1 Gross-Palmer formulation

Let V be a complex Hilbert space with Hermitian scalar product (·, ·)V anti-linear in the left variable. We denote A∗ the usual Hermitian adjoint of the linear operator A : V → V .

Definition 71 A real structure κ : V → V compatible with (·, ·)V is an anti-unitary involution, i.e a map such that, for all α, β ∈ C, v, w ∈ V , ¯ κ(αv + βw)=¯ακv + βκw, (κv, κw)V = (v, w)V , κκ =1.

Remark Since we will deal with Hilbert spaces we will always assume a real structure to be compatible with the scalar product and we will not explicitly say that it is compatible. For a real structure κ the following properties hold:

42 • Let Reκ v := (v + κv)/2, Imκ v := (v − κv)/(2i), then v = Reκ v + i Imκ v, κ Reκ = Reκ, κ Imκ = − Imκ, Reκ = Imκ i, Imκ = − Reκ i, so

V = Reκ V + i Imκ V = Reκ V + i Reκ iV,

which is a direct sum in the sense of vector spaces but it is not orthogonal with respect to (·, ·)V . • The form hhv, wiiκ := (κv, w)V is bilinear, symmetric and non-degenerate. • For any linear operator A : V → V we define its κ-transpose

Aκ := κA∗κ

κ which satisfies hhv, Awiiκ = hhA v, wiiκ. • If κ is another real structure on V , then letting U := κκ we have that U is a linear operator which is invertible and (Uv,Uw)V = (v, w)V , that is U is unitary. Note that V is a space of complex functions, then the map c : V → V given by taking the complex conjugate is a real structure.

Let ΓaV be the antisymmetric Fock space over the complex pre-Hilbert space V , i.e. the completion of the vector space ΛV with respect to the scalar product

(v1 ··· vn, w1 ··· wm)ΓaV = δn,m det (vj , wℓ)V , v1,...,wm ∈ V,n,m > 0. 16j,ℓ6n

Let CAR(V ) denote the algebra of canonical anti-commutation relations (CAR) over V , that is the unital, associative, complex, ∗-algebra generated by the identity and the generators c(v),a(v), v ∈ V which satisfy the following relations, for α, β ∈ C, v, w ∈ V : c(αv + βw)= αc(v)+ βc(w), c(v)= a(v)∗, and 0= {a(v),a(w)} = {c(v),c(w)}, {a(v),c(w)} = (v, w)V 1, where {a,b} stands for the anticommutator between a and b i.e. {a,b} = ab + ba. Note in particular that the creation operators c(v) are linear in v ∈ V , whereas the annihilation operators a(v) are anti-linear in v ∈ V . The algebra CAR(V ) has the following standard representation on ΓaV . Let (Cv)v∈V be the linear ∗ bounded operators in ΓaV acting on ΓaV as CvF = vF for any F ∈ ΛV and let Av = Cv be their adjoint. Then v 7→ Cv is a linear map, v 7→ Av is anti-linear and

{Cv, Cw} = {Av, Aw} =0, {Av, Cw} = (v, w)V , v,w ∈ V.

So (c(v), (v)) 7→ (Cv, Av), v ∈ V , is a representation of the CAR algebra CAR(V ) on ΓaV . Let Ω be the vacuum vector of ΓaV , i.e the unit element of ΛV and let ω : L(ΓaV ) → C the vacuum state defined as

ω(F )=(Ω, F Ω)ΓaV for F ∈ L(ΓaV ).

Recall that we constructed the centered complex Grassmann Gaussian X ∈ G(V ) associated to a com- plex Hilbert space V with real structure κ and covariance S : V → V via the CAR algebra (Cv, Av)v∈V as X(v)= CSv + Aκv, v ∈ V on the algebraic probability space (L(ΓaV ),ω). In particular let us recall that we have (cfr. (18))

ω(X(f1) ··· X(fn)) = Pf hhfi,Sfjiiκ, (87) 16i,j6n

We shall now collect some standard facts on Clifford C∗-algebras, and in particular their realization on the Fock space and the relation between the Fock vacuum and the unique tracial state on the Clifford algebra.

43 Definition 72 Let E be a complex vector space and hh·, ·ii a non-degenerate symmetric bilinear form on E. We denote by Cl(E, hh·, ·ii) the complex Clifford algebra on (E, hh·, ·ii), that is the unital, , over C, generated by the identity 1 and the symbols (γ(v))v∈E which satisfy {γ(v),γ(w)} = 2hhv, wii1, γ(αv + βw)= αγ(v)+ βγ(w), for all α, β ∈ C,v,w ∈ E.

Definition 73 Let V be a complex pre-Hilbert space with a real structure κ. We define a self-dual CAR algebra ClSD(V, κ) to be a complex Clifford algebra with respect to the symmetric bilinear form hh·, ·iiκ, where in addition the generators satisfy γ(v)∗ = γ(κv) for v ∈ V .

∗ 1/2 Proposition 74 If k·k denotes the (unique) C -norm of ClSD(V, κ), then kB(v)k = (v, v)V , where ∗ (B(v))v∈V are the generators of ClSD(V, U). The C -completion ClSD(V, κ) admits a unique trace, that is, a state τ : ClSD → C with the properties of being even (it is zero on the odd part of ClSD) and central (τ(AB)= τ(BA), A, B ∈ ClSD). Moreover we have, k ∈ N,

τ(B(v1) ··· B(v2k)) = Pf [sgn(ℓ − m)hhvℓ, vmiiκ], (88) 16ℓ,m62k and τ(B(v1) ··· B(v2k+1))=0, where Pf denotes the Pfaffian and sgn is the sign function. Proof The relation kB(v)k = (v, v)1/2, v ∈ V , is a consequence of the commutation relation {B(v)∗,B(v)} = {B(κv),B(v)} = (v, v) and the C∗-algebra property of the norm. It follows at once from the tracial 1 property of τ that τ(B(v)B(w)) = 2 τ({B(v),B(w)}) = hhv, wiiκ. The formula for the trace follows by induction. Indeed, by writing 1 τ(B(v ) ··· B(v )) = τ(B(v ) ··· B(v )+ B(v ) ··· B(v )B(v )), 1 2k 2 1 2k 2 2k 1 and noting that

2k j−1 ✟✟ B(v1) ··· B(v2k)+ B(v2) ··· B(v2k)B(v1)= (−1) {B(v1),B(vj )}B(v2) ···✟B(vj ) ··· B(v2k), j=2 X ✟✟ we get, using the induction hypothesis for τ B(v2) ···✟B(vj ) ··· B(vn) ,

2k  j−1 ✟✟ τ(B(v1) ··· B(v2k)) = (−1) {B(v1),B(vj )}τ B(v2) ···✟B(vj ) ··· B(v2k) j=2 X  2k j−1 = (−1) hhv1, vj iiκ Pf (sgn(m − ℓ)hhvℓ, vmiiκ) ℓ,m∈{2,...,6 j,...,2k} j=2 X = Pf (sgn(m − ℓ)hhvℓ, vmiiκ), 16m,ℓ62k where in the last line we used the extension rule for Pfaffians (cf. [39] formula (1.5)). Hence the trace τ is now uniquely defined for n = 2k even. By hypothesis the trace is normalized, hence τ(1) = 1, and even, i.e. τ(B(v1) ··· B(v2k+1))=0, k ∈ N (actually the fact of being even is implied by centrality when V is even dimensional). Therefore we have proved the formula in the statement and the uniqueness of τ. Finally, τ can be extended to the C∗-completion by continuity (cf. [134] Theorem 1.2.8). ✷

Definition 75 Assume that we have a complex pre-Hilbert space V with a real structure κ. Let CAR(V ) be the CAR algebra over V with generators Cv, Av, v ∈ V . We define the Gross-Palmer (G-P) fields BGP(v), v ∈ V , to be BGP(v) := Cv + Aκv, v ∈ V.

Remark 76 Let (B(v))v∈V , be the generators of ClSD(V, κ) then the ∗-homomorphism which extends B(v) 7→ BGP(v) to the whole ClSD(V, κ) is a realization of ClSD(V, κ) inside CAR(V ). Indeed, since κ2 =1, we have ∗ BGP(κf)= Cκf + Aκ2f = Cκf + Af = BGP(f) .

44 Moreover, since (κf, κg)= (f,g),

{BGP(f),BGP(g)} = {Cf , Aκg} + {Aκf , Cg}

= (κg,f)V + (κf,g)V = hhg,fiiκ + hhf,giiκ

= 2hhg,fiiκ

Hence (BGP(f))f∈V give a representation of ClSD(V, κ) within CAR(V ).

Proposition 77 Let (BGP(v))v∈V be as above. Then (BGP(v))v∈V is a ∗–representation of ClSD(V, κ) on ΓaV and the Fock vacuum state ω restricted to this representation coincides with the tracial state τ of ClSD(V, κ), namely

τ(B(v1) ··· B(vn)) = ω(BGP(v1) ··· BGP(vn)), v1,...,vn ∈ V,n ∈ N, (89) where (B(v))v∈V are the generators of ClSD(V, κ). Proof Since κ2 =1, we have

∗ ∗ ∗ BGP(v) = Cv + Aκv = Aκκv + Cκv = BGP(κv), v ∈ V.

Moreover,

{BGP(v),BGP(w)} = {Cv, Aκw} + {Aκv, Cw}

= (κw, v)V + (κv, w)V = hhw, viiκ + hhv, wiiκ

= 2hhv, wiiκ

Hence (BGP(v))v∈V give a representation of ClSD(V, κ) inside CAR(V ). A straightforward computation shows that the statement (89) holds for n =2. Then the result follows by noting that both states, τ and ω, are quasi-free (in the sense of Araki [14]), hence they are uniquely determined by their evaluation for n =2. ✷

A.2 A tracial formulation of Grassmann variables In this subsection we describe one way to relate the expectation of a complex Gaussian Grassmann random variable to the trace on a self-dual Clifford algebra. In the approach given in this subsection the construction of isotropic subspaces for certain bilinear forms reveals itself as crucial. One remarkable aspect of this approach is its “dual” nature with respect to the Osterwalder-Schrader approach. Perhaps its biggest disadvantage is that the random variables will be realized as elements of a Clifford algebra at the expenses of their anti-commutativity.

Assume we are given an orthogonal decomposition V = V+ ⊕ V− such that κV± = V∓. Then the subspaces V± are isotropic subspaces for the bilinear form hh·, ·iiκ, i.e. 0 = hhv, wiiκ if v, w ∈ V+ or v, w ∈ V−. That it is always possible to find such an isotropic decomposition is the subject of next proposition.

Proposition 78 Let κ be a real structure on V and assume V has even or infinite dimension. Then V admits an orthogonal decomposition V = V+ ⊕ V− such that κV± = V∓. Moreover, if c is another real structure on V then κ = Uc with U a unitary operator on V such that U c = U, where, as before, U c := cU ∗c, with U ∗ the adjoint of U. Finally, if c maps each V+, V− to itself, then, with respect to the decomposition V = W ⊕ W we have the block decomposition 0 u U = , (90) uc 0   for some unitary operator u : W → W . Conversely if U and c are as above for any (complex linear) unitary u : W → W , then κ := Uc is a real structure on V compatible with (·, ·)V .

45 Proof Let (fj )j∈N be an orthonormal basis of V . If fj + κfj 6= 0, let ej := fj + κfj , otherwise let + − ej := i(fj −κfj), then (ej )j∈N, is an orthogonal basis of V invariant under κ. Let gj := e2j−1 +ie2j,gj := + − ∼ e2j−1 − ie2j , j ∈ N. Then (gj )j∈N, (gj )j∈N generate orthogonal isomorphic subspaces V+ = V−. Let + − π+, π− be the respective projections. Then, since κ : gj 7→ gj , we have κπ+ = π−κ. Hence V+, V− are isotropic. Let c be another real structure on V . Let U := κc, then U is linear and unitary since (Uv,Uw)V = ∗ 2 ∗ c (κcv,κcw)V = (v, w)V . Moreover UU =1= κ = UcUc gives that cUc= U , that is U = U . If cV± = V± then both π+, π− commute with c. Because of this and the fact that V+, V− are isotropic we have U = π+Aπ− + π−Bπ− for some (complex linear) operators A, B. Let us identify V+ =∼ V− and denote them both by W , then A, B : W → W . Because U is unitary A and B are unitary. Finally, from U = U c, we get AcBc=1. Hence if we let u := A we get the decomposition (90). The last statement follows by direct verification that, for any unitary operator u : W → W , κ = Uc, 2 U, c as stated, we have κ = I, κi= −iκ, and (κv, κw)V = (w, v)V so κ is another real structure. ✷ Following Gross (cf. Lemma 3.1 in [70]) we give the following definition.

Definition 79 Let V = V+ ⊕ V− be a complex vector space which decomposes into a direct sum of isotropic subspaces with respect to a symmetric, non-degenerate, bilinear form hh·, ·ii : V × V → C. We define the vector space isomorphism θ :ΛV → Cl(V, hh·, ·ii) by i. θ(1) = 1, θ(v)= B(v), v ∈ V ;

ii. θ(v1 ∧···∧ vn ∧ w1 ∧···∧ wm)= B(v1) ··· B(vn)B(w1) ··· B(wm),

if vj ∈ V+, wℓ ∈ V , or wℓ ∈ V−, vj ∈ V , j =1, . . . , .n, ℓ =1,...,m.

Remark 80 The map θ order the factors so that the generators belonging to V+ are all on the left, in essence is a Wick ordering map with respect to the decomposition V = V+ ⊕ V−. Note that θ : ΛV → Cl(V, hh·, ·ii) cannot be extended to an algebra isomorphism since the exterior algebra ΛV and the complex Clifford algebra Cl(V, hh·, ·ii) are not isomorphic. By the isotropy of the symmetric bilinear form hh·, ·ii, the map θ is an algebra homomorphism when restricted to V+ or V−. Finally note that θ has the following nice interpretation in the language of Hopf algebras (cf. Section 2.2). Since the restrictions θ+, respectively θ−, of θ restricted to ΛV+, respectively ΛV−, are algebra isomorphisms, we let ϑ :ΛV+ ⊗ ΛV− → Cl(V, hh·, ·ii) ⊗ Cl(V, hh·, ·ii) be the algebra isomorphism (with its image) defined by ϑ := θ+ ⊗ θ−. Moreover let mCl : Cl(V, hh·, ·ii) ⊗ Cl(V, hh·, ·ii) → Cl(V, hh·, ·ii) denote the multiplication in Cl(V, hh·, ·ii). Then we have

θ(F )= mCl[ϑ(ΠΛV+ ⊗ ΠΛV− )∆F ], F ∈ ΛV, (91)

where ∆ denotes the co-product of ΛV . To see (91) we first note that (ΠΛV+ ⊗ ΠΛV− )∆ it is nothing but the isomorphism Λ(V+ ⊕ V−) =∼ ΛV+ ⊗ ΛV−. Now, this isomorphism sends F → F+ ⊗ F−, F+ ∈ ΛV+, F− ∈ ΛV−. Finally by the definition of θ, we see that θ applied to an element F = F+F− ∈ ΛV , gives θ(F+F−)= θ(F+)θ(F−)= mClϑ(F+ ⊗ F−). Hence (91) is verified.

Remark 81 Let κ be a real structure on V and let V = V+ ⊕ V− be an orthogonal, hh·, ·iiκ-isotropic decomposition. With respect to this decomposition, let c := σκ with

0 1 σ = . 1 0  

Then c is a real structure on V and it preserves the decomposition V+ ⊕V− . If κ is another real structure on V which with respect to the decomposition V+ ⊕ V− has the form

0 u κ = κ c u 0   for some unitary u : V+ → V+, then u 0 κ = κ κ. 0 u  

46 If S : V → V is unitary, κ-antisymmetric, and preserves the decomposition V+ ⊕ V− then S has the form

v 0 S = κ 0 −v   for some unitary v : V+ → V+.

Theorem 82 Let X ∈ G(V ) be a complex Gaussian Grassmann variable with covariance associated to a triple (V, κ,S), where κ is a real structure on V and S = −Sκ is the covariance of X. Assume S is unitary and assume there exists an orthogonal, hh·, ·iiκ-isotropic decomposition V = V+ ⊕ V− which is preserved by S : V± → V±. Let π+, π− denote the projections onto respectively V+ and V−. Then

κ := S(π+ − π−)κ is a real structure on V with respect to which, if v±, w± ∈ V± and v, w ∈ V we have,

hhv±, wiiκ = (κv±, w)V = ±(κSv±, w)V = ±hhSv±, wiiκ (92)

hhv, w−iiκ = hhw−, viiκ = −hhSw−, viiκ = hhw−, Sviiκ = hhSv,w−iiκ. (93)

As a consequence the map θ :ΛV → ClSD(W, κ) in Definition 79 satisfies

ω(X(F )) = τ(θ(F )), F ∈ ΛV. (94)

Proof By Proposition 78 we see that κ is a real structure on V and that equations (92) and (93) are satisfied. Hence, from the definition of θ, τ, and ω, we get

τ(θ(B(v)B(w))) = ω(X(v)X(w)), v,w ∈ V, which proves (94) for n = 2. To prove the general case, let us first note that without loss of generality we can take S = −π+ + π−.

Moreover, from (91) in Remark 80, it is enough to consider elements of the form F = F+F− ∈ ΛV , + + F+ ∈ ΛV+, F− ∈ ΛV−. Finally, by linearity we can restrict our attention to F = F+F− = v1 ∧···∧ vk ∧ − − vk+1 ∧···∧ vn . Now, by the definition of θ, we have

+ + − − + + − − θ(v1 ∧···∧ vk ∧ vk+1 ∧···∧ vn )= B(v1 ) ··· B(vk )B(vk+1) ··· B(vn ). Hence, by comparing (88) with (87), we see that to prove (94) it will suffice to show that

# # # # Pf [sgn(ℓ − m)hhvℓ , vmiiκ]= Pf hhvℓ , Svmiiκ, 16ℓ,m6n 16ℓ,m6n

# + # − where vj = vj for j =1,...,k and vj = vj for j > k. Note first that if n 6=2k then both terms are zero because the matrices inside the Pfaffians would be degenerate. Finally observe that, since S = −π+ + π− and since V+, V− are isotropic, we have

+ + + + + − + − 0= hhvℓ , vmii = hhvℓ , Svmii, hhvℓ , vk+mii = hhvℓ , Svk+mii, 1 6 ℓ,m 6 k, and − + − + − − − − hhvk+ℓ, vmii = −hhvk+ℓ, Svmii, 0= hhvk+ℓ, vk+mii = hhvk+ℓ, Svk+mii, 1 6 ℓ,m 6 k. Hence the matrices inside the Pfaffians are the same. This concludes the proof. ✷ 2 2 2 −1/2 2 With notation as in Section 5.1, let now W = L (R ) ⊗ C , u := T ∇/ + mf and T := mf − ∆, with T as self-adjoint operator on L2(R2) ⊗ C2 with domain W := S(R2) ⊗ C2. Note that T is positive  essentially self-adjoint and u is unitary. Take V := W ⊕ W with π+, π− the projections onto the two copies of W and consider the scalar product

−1/2 −1/2 (v, w)V := (π+v,T π+w)L2(R2)⊗C2 + (π−v,T π−w)L2(R2)⊗C2 , v,w ∈ V.

47 We denote by c the operation of complex conjugation of a function in L2(R2)⊗C4. Note that T commutes with c, hence c is a real structure on V . If we let κ := Uc, with

0 u∗ U := , cuc 0   by Proposition 78 and the fact that both U and c commute with T , we have that U is another real structure on V . If we take S : V → V given by

1 0 S = , (95) 0 −1   in the decomposition V = W ⊕ W , then we have for all v, w ∈ V ,

0 u∗ (κv,Sw) = cv,T −1/2Sw V cuc 0   L2(R2)⊗C2 0 u = cv,T −1/2 w −cu∗c 0    L2(R2)⊗C2 ˜ ˜ = (cv,Kw)L2(R2)⊗C2 = ω(ψ(v)ψ(w))

2 2 2 2 2 2 with K : L (R ) ⊗ C → L (R ) ⊗ C as given in Section 5.1 and (ψ˜(v))v∈V the complex Grassmann Gaussian introduced there to represent the Euclidean Dirac field. As a consequence we have:

Corollary 83 On the pre-Hilbert space V = S(R2) ⊗ C4 there exists a real structure κ and orthogonal projectors π+, π− such that π+V =∼ π−V and, for fj ,gj ∈ V and n ∈ N,

ω(ψ˜(π+f1) ··· ψ˜(π+fn)ψ˜(π−g1) ··· ψ˜(π−gn)) = τ(B(π+f1) ··· B(π+fn)B(π−g1) ··· B(π−gn)), (96) where (B(v))v∈V , are the generators of ClSD(V, U) and τ is its trace and where (ψ˜(v))v∈V is the Gaussian Grassmann field of Section 5.1. Proof This is a consequence of Theorem 82 once we show (96) for n = 2. By direct computation, for all v, w ∈ V we have indeed

−1/2 τ(B(π+v)B(π−w)) = (κπ+v, π−w)V = (cπ+v,T uπ−w)L2(R2)⊗C2 ˜ ˜ = (cπ+v,Kπ−w)L2(R2)⊗C2 = ω(ψ(π+v)ψ(π−w)). ✷

A.3 The real-wave representation of Grassmann Gaussians Theorem 82 gives a realization of the Grassmann Gaussian variables in terms of the trace state on a Clifford algebra. This realization is obtained through the vector-space map θ, which can be interpreted as a sort of “Wick-ordering” with respect to a suitable decomposition V = V+ ⊕ V−. Since the map θ is not an algebra isomorphism that realization is not a representation. Here we are going to give an actual representation of Grassmann Gaussian variables as multiplication operators on a non-commutative L2 space in full analogy to the case of the usual (commutative) Gaussian variables. Moreover the represen- tation we give here will not depend on any decomposition V = V+ ⊕ V−, hence making the construction intrinsically more natural.

Let K =ΓaV be the antisymmetric Fock space over the complex Hilbert space V with a real structure κ. Following the exposition of Gross [68] one considers the Clifford algebra C0 ⊆ L(K) generated alge- braically by the operators (BGP(v))v∈V . Let C be the von Neumann algebra generated by C0 and consider on C the state tr(u)= ω(u)= hΩ,uΩiK for u ∈ C. This state is tracial (on C), that is tr(CB) = tr(BC) for all B, C ∈ C and faithful, i.e. tr(A∗A)=0 ⇒ A =0. So the scalar product (A, B) 7→ tr(A∗B) is posi- tive definite and the closure of C with respect to the corresponding norm defines the Hilbert space L2(C).

48 By a theorem of Segal ([154], but see also Theorem 5 in [68]) there is an isomorphism D : L2(C) → K such that u ∈ C ⊆ L2(C) is sent to uΩ ∈ K. Let κ be, as before, the real structure on V . By the proof of Proposition 78 we know that there exists a κ-invariant orthonormal basis (ej )j of V . Then the −1 −1 inverse D : K → C satisfies D (ei1 ∧···∧ ein ) = BGP(ei1 ) ··· BGP(ein ). Note however that D is a canonical isomorphism and in particular (unlike the map θ of the previous section) does not depend on any particular decomposition V = V+ ⊕ V−. By Corollary 5.2 of [68], writing La (resp. Ra) for the left (resp. right) multiplication of a ∈ C on L2(C), one has

−1 −1 DLBv D = Cv + Aκvβ, DRBv D = (Cv − Aκv)β, v ∈ V (97) where β = Γ(−1) : K → K is the unitary map corresponding to the second quantization of the unitary operator v 7→ −v on V . Let us denote by βˆ = D−1βD the pullback on L2(C) of this unitary.

Remark 84 Recall that the standard centered Gaussian measure µ corresponding to an Hilbert space 2 V supports the Itô–Wiener isomorphism L (µ) ≈ ΓsV , where ΓsV is the symmetric Fock space over V . The gaussian random variable X with law µ is then realized as an unbounded self-adjoint operator acting in L2(µ) given by multiplication. The space L2(C) can be considered to be the fermionic equivalent of this construction, in that it provides the representation of the antisymmetric Fock space as a space of “L2 functions” over a non-commutative probability space (C, tr). The von-Neumann algebra C can be interpreted as L∞(C) acting on L2(C) and fermionic random variables can be constructed, according to (97), as (left and right) multiplication operators on L2(C). We want now to understand this construction in the context of our Grassmann Gaussian variables. This will allow us to replace the “particle” Fock space K by the space L2(C) of “wave-functions” for the von Neumann Clifford algebra C. While this is not important in the theory exposed in our work, it could play a role as soon as we have to deal with unbounded operators, for which the Lp spaces of unbounded operators associated to the non-commutative probability space (tr, C) could become useful [68].

As before we consider an operator S : V → V such that Sκ = −S. The triple (V, κ,S) determines a complex Grassmann Gaussian variable X ∈ G(V ) with covariance hh·,S · iiκ (cf. Definition 17). By the Osterwalder–Schrader construction, this Gaussian has been obtained via the concrete representation as sum of creation and destruction operators over the antisymmetric Fock space ΓaV with the state ω given by the Fock vacuum and as in the previous subsection we have

X(v)= CSv + Aκv, v ∈ V.

By Segal’s duality D there will be multiplication operators on L2(C) corresponding to our Grassmann Gaussian. Let us work out their precise form: by (97) we have

ˆ −1 ˆ −1 D(LBv + RBv β)D =2Cv,D(LBv − RBv β)D =2Aκv, v ∈ V.

Therefore for any v ∈ V , 1 1 Xˆ(v) := D−1X(v)D = [L + L ]+ [R − R ]βˆ 2 BSv Bv 2 BSv Bv = L − R β,ˆ BQ+v BQ−v where Q± = (1 ± S)/2. By construction, (Xˆ(v))v∈V are the generators of a Grassmann algebra and we can extend them to an homomorphism Xˆ :ΛV → L∞(C) such that

tr(Xˆ(F )1) = ω(X(F )), for all F ∈ ΛV . This shows that, if needed, we can realize Grassmann Gaussians over the L2-space over the non-commutative measure space (C, tr) in full analogy with the bosonic setting. Note that in the particular case of the Euclidean Dirac field the operator S has the diagonal form (95) and therefore the operators Q± = π± are the projections onto the isotropic decomposition V = V+ ⊕ V− corresponding to the two fields ψ and ψ¯ which constitute the Euclidean Dirac field ψ˜ (cf. Section 5.1).

49 B Convergence of the perturbative series: finite dimensional case

In this appendix we sketch some implications of the Pauli exclusion principle for the existence of global solutions to finite-dimensional Grassmann SDEs. In principle some of these considerations also apply to some more realistic models like the Yukawa2 model however, to our surprise, stochastic quantization, at least in the perturbative regime, can be carried on without establishing them, as we demonstrated in Section 5.

The solutions to the non-linear equation (43) can be represented via a series. We investigate now the properties of this series and its absolute convergence. The explicit control of the series is not needed for the development of the theory of stochastic quantization as we have shown. For simplicity let us assume that (vα)α=1,...,N is a finite basis of V and that F ∈ Hom(V, ΛV ) is given by a sum of cubic monomials

α F (vα)= λα1α2α3 ψ(vα1 )ψ(vα2 )ψ(vα2 ), α ,α ,α 1X2 3 α where ψ ∈ Hom(V, ΛV ) is the canonical injection of V into ΛV and (λα1α2α3 ) a family of coefficients in R. Moreover we take A = −I. Equation (43) has then the integral formulation given by

t −(t−s) Ψt(v)=Φt(v)+ Ψs(e F (v))ds, t > 0, v ∈ V, (98) Z0 where −t A A −t Φt(v) := Ψ0(e v)+ Bt (v) − B0 (e v), t > 0, v ∈ V.

By iteratively expanding Ψs on the right hand side of (98) we obtain a series expansion for Ψt of the form Ψt = Jτ (Φ)(t) (99) τ X = J•(Φ)(t)+ J[•••](Φ)(t)+ J[[•••]••](Φ)(t)+ ··· + J[[•••][•[••[•••]]•]•](Φ)(t)+ ··· The series is indexed by (planar) trees τ which have branches of order 3 and where J is a multilinear integral operator such that α J•(Φ)(t) =Φt(vα) t α −(t−s) α α1 α2 α3 J[τ1τ2τ3](Φ)(t) = e λα1,α2,α3 Jτ1 (Φ)(s) Jτ2 (Φ)(s) Jτ3 (Φ)(s) ds α ,α ,α 0 1X2 3 Z where • denotes the simple tree and [τ1,...,τ3] the tree with branches τ1,...,τ3. Our goal is to prove that the above series converges for all times and derive estimates for the norm kΨtk of the solution Ψ. Expanding the expression for Jτ (Φ) we have

αp αp Jτ (Φ) = λα α α Φ (sp− ) dsp  p1 p2 p3      Z p∈YI(τ) p∈YL(τ) p∈YI(τ)       where I(τ) denotes the internal nodes, L(τ) the leaves, p the parent node and p1,p2,p3 the three children α of every internal node and Φ (t)=Φt(vα). We have |τ| = |L(τ)| + |I(τ)|, |τ| =1+3|I(τ)|, |L(τ)| = 1+3|I(τ)| − |I(τ)| =1+2|I(τ)|.

α 2 α α α Note that (Φt ) =0 and that the increments Φ (t,s) := Φt − Φs have the norm bound

t kΦα(t,s)k2 6 e−(t−r)dr + |e−(t−s) − 1|2 6 2(1 − e−(t−s)) =: H(t − s) Zs which can be estimated as kΦα(t,s)k 6 |t − s|1/2. α 2 The following Lemma uses (Φt ) =0 (Pauli exclusion principle) to derive good estimates for products of fields.

50 Lemma 85 Assume m > 1. For any n > 1 and any t1,...,tn ∈ [0,T ], α1,...,αn ∈{1,...,N} we have

Cn+1T n/2 kΦα1 ··· Φαn k 6 t1 tn (n!)1/2 where C is a universal constant depending only on N.

Proof Partition [0,T ] in r = n/(2N) equal intervals (Ik)k and let (sk)k be the centers of those intervals. αi αi αi αi Now in the product replace each Φt1 by Φt1 = Φsk +Φ (ti,sk) where sk is the nearest to ti of the α1 αn n αi centers. By doing so we rewrite Φt1 ··· Φtn as a sum S of 2 products of fields which are either Φsk or αi αi Φ (ti,sk), moreover for a given Ik there cannot be more than N factors Φsk by the exclusion principle. αi Denote by nk the number of (ti)i in Ik. Then n = k nk and using the estimate kΦ (ti,sk)k 6 |ti −sk| 6 |Ik| = (2NT )/n and the fact that in each cell we have at most nk − N increments, we have P r 2NT n−Nr 2NT n/2 Cn+1T n/2 kSk 6 2n |I |nk−N 6 2n =2n 6 . k n n (n!)1/2 kY=1     ✷

Using this lemma above we can estimate

αp αp kJτ (Φ)(t)k 6 λα α α Φs dsp  p1 p2 p3  p−   Z p∈I(τ) p∈L(τ) p∈I(τ) Y Y Y    

C|L(τ) |+1t|L(τ)|/2 . |λ||I(τ)|N |L(τ)| (|L(τ)|!)1/2 where we estimated all the integrals by constants uniformly in t due to the presence of the exponential factors. The number of trees τ with a given number of nodes |τ| = n is no more than Dn for some n and D > 0 and |L(τ)| =1+2|I(τ)| =1+(2/3)(|τ|− 1)= 1/3+2|τ|/3 so we obtain at the end for the solution Ψt of (98)

C2n/3tn/3+1/6 kΨ k 6 kJ (Φ)(t)k . Dn|λ|(2/3)(n−1)N 2n/3 (100) t τ (n/3+1/6)! τ X nX>0 which is a series which converges for all t (with some stretched exponential behavior).

Theorem 86 There exists an increasing function E(t) depending on N, |λ|,m such that for all m> 0,

sup kΨsk 6 E(t), s6t where recall that (Ψt)t>0 is the unique solution of (98). In particular this shows again that explosion is not possible for this kind of SDEs, unlike their bosonic analogs.

Our goal now is to remove the time dependence on our estimate for the solution exploiting the exponential decay. Let us partition the interval [0, ∞] into intervals (Ih)h>0 of size 1. In each of these intervals we will use the finer partition of the previous lemma to estimate products, which will then not depend on T anymore. Of course now the problem is that we have a bound of the form

Cnh+1 α1 αn 6 kΦ (t1) ··· Φ (tn)k 1/2 (nh!) hY>0 where nh is the number of time variables in Ih and h>0 nh = n, and so far there is no useful upper bound on this (since we observe that we can have the situation where nh = 1 for all h and so we loose P the factorial contribution). Let Q be the number of intervals with nh > 0. In the expression for Jτ (Φ)(t)

51 ′ there are at least Q factors of the form e−(s−s ), s′

Cnh+1 Cn|λ|(2/3)(n−1) 6 (2/3)(n−1) n −Q/2 6 −Q/2 n/2 kJτ (Φ)(t)k |λ| N e 1/2 1/2 e Q (nh!) (n!) QX>1 h:Ynh>0 QX>1 where we used that 1 1 n! 1 6 = Qn. nh! n! k1! ··· kQ! n! h:Ynh>0 k1,...,kXQ > 1 k1 + ··· + kQ = n −Q/2 n/2 n 1/2 Now, using the bound Q>1 e Q . c (n!) we obtain that

(2/3)(n−1) n P kJτ (φ)(t)k . |λ| C (101) so provided |λ| is small enough (depending on our choice of N) we conclude, from the representation (99), that the uniform bound supt>0 kΨtk . 1 holds.

C Besov spaces of Banach algebras

In this section we want to recall some results about Besov spaces of functions (or distributions) from Rd taking values in a Banach algebra A. All the results of this section can be found in [7, 8] for the theory of Besov spaces taking values in a Banach space and [169] for weighted Besov spaces. We present here only the Rd case. The definition of Besov spaces on Td is similar, and all the results announced here also hold on Td.

We denote by S(Rd) the set of smooth functions f ∈ S(Rd) such that

ℓ α kfkℓ,α := k(1 + |x|) |D f|kL∞(Rd) < +∞

d ′ d d where ℓ ∈ R+ and α ∈ N . We denote by S (R ) the strong dual of S(R ) (equipped with the topology d d of the semi-norms k·kℓ,α) and by OM (R ) the set of smooth functions f ∈ OM (R ) such that for any d α ∈ N there exists ℓα ∈ R+ for which

−ℓα α k(1 + |x|) |D f|kL∞(Rd) < +∞. We use the notations

d d ′ d ′ d d d S(R , A)= S(R )⊗Aˆ , S (R , A)= S (R )⊗Aˆ , OM (R , A)= OM (R )⊗Aˆ , where A is a Banach algebra and A⊗ˆB is the completion with respect to the natural metric of the algebraic tensor product A ⊗ B when A is a Fréchet nuclear space and B is a Banach space (see Remark ). Note that these tensor products are well defined since the spaces S(Rd),S′(Rd), O(Rd) are nuclear. It is important to note that d d d d ′ d (S(R , A) ֒→ OM (R , A) ֒→ S (R , A where the arrows mean that each space is continuously embedded and dense in the following one.

Theorem 87 It is possible to define uniquely h·, ·i : S(Rd, A)×S′(Rd, A) → A, · : S(Rd, A)×S′(Rd, A) → ′ d d ′ d ′ d d ′ d d α ′ d S (R , A), · : S(R )×S (R , A) → S (R , A), ∗ : S(R )×S (R , A) → OM (R , A) and D : S (R , A) → S′(Rd, A) (where α ∈ Nd) which extend in an epicontinuous way the following operations: any f ∈ S(Rd), ′ d u ∈ S (R ) and a1,a2 ∈ A we have

hf ⊗ a1,u ⊗ a2i = hf,uia1a2 (f ⊗ a1) · (u ⊗ a2) = (fu) ⊗ (a1a2) f · (u ⊗ a2) = (fu) ⊗ a1 f∗(u ⊗ a2) = (f∗u) ⊗ a1 α α D (u ⊗ a1) = (D u) ⊗ a1 where hf,ui is the normal pairing in S(Rd) × S′(Rd), (fu) is the product in S(Rd) × S′(Rd), (f∗u) is the convolution in S(Rd) × S′(Rd) and Dα is the α derivative in S′(Rd).

52 Proof The proof of this theorem can be found in [8] Appendix 1. ✷ We recall the definition of Littlewood–Paley blocks: let χ, ϕ be smooth non-negative functions from Rn into R such that

• supp(χ) ⊂ B 4 (0) and supp(ϕ) ⊂ B 8 (0) \ B 3 (0), 3 3 4

−j n • χ, ϕ 6 1 and χ(y)+ j≥0 ϕ(2 y)=1 for any y ∈ R , • supp(χ) ∩ supp(ϕ(2−Pi·)) = ∅ for i > 1, • supp(ϕ(2−j ·)) ∩ supp(ϕ(2−i·)) = ∅ if |i − j| > 1,

n where by Br(x) we denote the ball of center x ∈ R and radius r > 0. We introduce the following −j 2 −ℓ/2 notations: ϕ−1 = χ, ϕj (·)= ϕ(2 ·), Kj =ϕ ˆj and, for any ℓ> 0, we define ρℓ(x) := (1 + |x| ) .

Remark 88 It is always possible to choose χ and ϕ in such a way that there exists some constants γ−1,γ > 0 and 0 <θ−1,θ< 1 such that

θ−1 θ |K−1(y)| . exp(−γ−1|y| ) and |K0(y)| . exp(−γ|y| ),

(see, i.e., [147] Section 1.2.2 Proposition 1 or [120]).

If v ∈ S′(Rd, A) and if i ∈ Z, i > −1 we define i-th Littlewood–Paley block as follows

d ∆iv = Ki∗v ∈ OM (R , A).

Then, if s ∈ R, p, q ∈ [1, +∞] and ℓ ∈ R we define the function

1/q +∞ jsq q s kvkB (Rd,A) = 2 k∆j vkLp(Rd,A) , p,q,ℓ  ℓ  j=−1 X   js when q ∈ [1, +∞) and kvk s Rd = sup (2 k∆jvk p Rd ), where k·k p Rd is the norm in the Bp,+∞,ℓ( ,A) j Lℓ ( ,A) Lℓ ( ,A) p d space Lℓ (R , A) that is 1/p p kfk p Rd = (kf(y)kAρℓ(y)) dy , Lℓ ( ,A) Rn Z  if p ∈ [1, +∞) and kfk ∞ Rd = sup (kf(y)kAρℓ(y)), Lℓ ( ,A) y∈Rd

d s d if p =+∞. For any v ∈ S(R , A) the norm kvk s Rd < +∞ is finite. Then we look at B (R , A) Bp,q,ℓ( ,A) p,q,ℓ d ′ d as the closure of S(R , A) in S (R , A) with respect to the norm k·k s Rd . Hereafter, if s ∈ R, Bp,q,ℓ( ,A) s d s d s d p, q ∈ [1, +∞] and ℓ ∈ R, we use the following notations Cℓ (R , A) := B∞,∞,ℓ(R , A), Bp,q(R , A) = s d s d s d s s d Bp,q,0(R , A), C (R , A) := B∞,∞,0(R , A), Bp,q,ℓ := Bp,q,ℓ(R , R) etc.

In this paper we need only the next three results.

s d s d Theorem 89 For any s,ℓ ∈ R and v ∈Cℓ (R , A) we have that ρℓv ∈C (R , A) and

kvk s Rd ∼kρ vk s Rd . Cℓ ( ,A) ℓ C ( ,A)

Proof The estimate holds by Theorem 6.5 in [169] for weighted Besov spaces on Rd with polynomial-like weights. The theorem for a general Banach algebra A is a straightforward generalization. ✷

Theorem 90 Consider m> 0, α,s,ℓ ∈ R such that s,s + α 6∈ N then we have that (−∆+ m)−α, where d s d s+α d ∆ is the standard Laplacian on R , is a continuous linear map from Cℓ (R , A) into Cℓ (R , A)

53 Proof This is exactly Theorem 5.3.2 of [8] for the unweighted case ℓ = 0. The theorem can be easily extended using the techniques of Chapter 6 of [169] to Besov spaces with weights ρℓ. ✷

α Theorem 91 Let α > 0 and consider a ∈ B1,∞, then we have that the convolution a∗ can be extended in a unique continuous way from an operator from S(Rd, A) into itself, into an operator from Cs(Rd, A) into Cs+α(Rd, A).

α α/2 Proof Using the notations of [7], if a ∈B1,∞ then ((−∆+1) (a)) ∈FL1, then the thesis follows from Corollary 6.4 in [7]. ✷

d d d Theorem 92 Let s> 0 and ℓ1,ℓ2 ∈ R we have that the product · : S(R , A) × S(R , A) → S(R , A) can s Rd s Rd s Rd be (uniquely) extended in a continuous way from Cℓ1 ( , A) ×C ( , A) into Cℓ1+ℓ2 ( , A), furthermore s Rd s Rd for any v1 ∈Cℓ1 ( , A) and v2 ∈Cℓ2 ( , A) we have

kv1 · v2kCs (Rd,A) . kv1kCs (Rd,A)kv2kCs (Rd,A), ℓ1+ℓ2 ℓ1 ℓ2 Proof The proof can be found in [8] in the unweighted case. The proof in the Besov spaces with polynomial weights ρℓ is similar and can be done using the techniques of Chapter 6 of [169]. ✷

D Convergence of the perturbative series: infinite dimensional case

In this section we want to consider the stationary solution to the equation

t 2 ˜ s 2 (∆−mf )(t−τ) ˜ ˜A Ψε,λ2,t(x)= λ e (Fε,Y (Ψτ (·)))dτ + Bε,t (102) Z−∞ 2 ˜ ˜ as a function of λ , i.e. we want to study the regularity of the function λ 7−→ Z ˜A (λ) from a Fε,Y ,Bε,t neighborhood U ⊂ R of 0 into C0(R, Cq(R2, A⊗ C2)) (here 0

˜ Theorem 93 Under the Hypotheses of Theorem 59, the function Z ˜A (λ) is real analytic (see Defi- Fε,Y ,Bt nition 101) in a neighborhood of 0. An easy consequence of Theorem 93 is the following corollary on the Schwinger functions of the (regularized) Yukawa2 model.

Corollary 94 Under the Hypotheses of Theorem 59, we have that the Schwinger functions of the (regu- ∞ 2 4 larized) Yukawa2 model, i.e. for any f1,...,fn ∈ Cc (R ) ⊗ C

V ω ψ˜(f1) ··· ψ˜(fk)e ε,hn ˜ s ˜ s Sn(f1,...,fn|λ) = lim = ω(Ψ 2 (f1) ··· Ψ 2 (fk)) (103) Vε,h ε,λ ,t ε,λ ,t hn→1  ω (e n ) 

(where hn is a sequence of spatial cut-off as defined in Section 5) is a real analytic function of the coupling constant λ.

q 2 Proof We have that the functional (Y1,...,Yn) 7→ ω(Y1(f1) ··· Yn(fn)) (where (Y1,...,Yn ∈C (R , A⊗ C4)) is a multilinear continuous functional (and so real analytic), from Cq(R2, A⊗ C4) into C (since ω is continuous and linear from A into C and the product in A is also continuous and multilinear). Since the composition of real analytic functionals is real analytic the thesis follows. ✷

54 Remark 95 Theorem 93 and Corollary 94 do not depend on the fact that the non-linearity Fε,h,Y has the specific form of the case of Yukawa interaction (i.e., for example, that is a third degree polynomial or on the specific non local form). In particular Theorem 93 and Corollary 94 extend easily to the (regularized) quantum field fermionic models mentioned in Appendix E. The rest of this appendix contains the proof of Theorem 93. Theorem 93 can be seen as an infinite dimensional generalization of the results in Appendix B, where, between the other things, it is proved that the real analyticity of Theorem 93 holds for finite dimensional systems. In our case we can consider the finite dimensional approximation

t h ,s 2 h ,s Ψ˜ R = λ˜ e(∆−mf )(t−τ)P (F (Ψ˜ R ))dτ + B˜A (v) (104) ε,λ,N,R,t˜ N ε,hR,Y ε,R,τ ε,R,t Z−∞ ˜ hR,s and the relative function Z ˜A (λ) := Ψ˜ . PN (Fε,hR,Y ),Bε,R,t ε,λ,N,R,t˜

˜ ˜ Lemma 96 The function λ 7→ Z ˜ A (λ) is real analytic in a neighborhood of 0 PN (Fε,hR,Y ),Bε,R,t Proof The result follows from the proof of convergence of perturbation series given in Appendix B (in particular from inequalities (100) and (101)) for finite dimensional Grassmannian equation and from the ∞ 2 4 ∞ 2 fact that the topology on G(PN (C (TR, A⊗ C ))) does depend on the norm chosen on PN (C (TR, A⊗ 4 ∞ 2 4 ✷ C )) since all the norms on PN (C (TR, A⊗ C )) are equivalent being, this space, finete dimensional. Instead of proving the Theorem 93 by trying to generalize the computation done in Appendix B we present a proof using the approximation method described in Section 5, since it is closer to the SPDE approach followed in the main part of the present paper. For this reason we want to introduce some sort of apriori estimate characterizing real analytic functions. This is done exploiting the notion of formal power series and majorants method (see, e.g., [170] for an introduction to the subject). We call a formal power series around 0 taking values in the Banach space X the following object:

+∞ f = f (n)λn, λ ∈ C, (105) n=0 X where f (n) ∈ X. We write f(0) = f (0) and also

+∞ n−1 dnf = f (n)λn−k (k − i). dλn i=0 kX=n Y When the power series (105) is absolutely convergent for 0 < |λ| < ε (for some ε > 0) we identify the +∞ (n) n formal series (105) with the real analytic function λ 7→ f(λ) := n=0 f λ ∈ C defined for |λ| <ε. P Definition 97 Let fX be a formal power series around 0 taking values in the Banach space X and let K be a formal power series around 0 taking values in C. We say that f is majored by K, and we write fX P K, if for any n ∈ N we have

dnf dnK X (0) = n!kf (n)k 6 (0) = n!K(n), dλn X dλn X

+∞ (n) n where K(λ)= n=0 K λ . P Remark 98 If fC is a formal power series in C and K is an analytic function such that fC P K then also fC is an analytic function. Furthermore if the power series defining g (in a neighborhood of 0) converges absolutely for λ = R> 0, then the power series defining fC converges absolutely for λ ∈ C with |λ| 6 R and furthermore sup |fC(λ)| 6 K(R). λ∈C,|λ|6R

55 In general it is possible to define the sum and product of formal power series. Furthermore if g : C → C is an entire function and fC is a formal power series taking values in C it is possible to define the composition g ◦ fC as a formal power series taking values in 0. +∞ (n) n (n) Let g be an analytic function such that 0 P g, i.e. g(λ) := n=0 g λ such that g > 0, consider c > 0 and define P Gg,c(a, λ)= λg(a)+ c.

Lemma 99 For any entire function g such that 0 P g and c > 0 there exists a unique analytic function K : U → R (where U is a suitable neighborhood of 0) such that

K(λ)= Gg,c(K(λ), λ) (106) for any λ ∈ U. Furthermore if fC is a formal power series such that fC P Gg,c(fC(·), ·) then fC P K. Proof Equation (106) admits a unique real analytic solution K(λ) defined in a neighborhood U of 0 and taking values in a neighborhood V of c ∈ R. Indeed we have that the point (c, 0)=(k, λ) is solution to the equation k − Gg,c(k, λ)=0. (107) Furthermore we have that

′ ∂k(k − Gg,c(k, λ))|(k,λ)=(c,0) = (1 − λg (k))|(k,λ)=(c,0) =1.

Thus by the theorem for holomorphic functions (see, e.g., [113] Chapter 3, Theorem 3.11), there is are two sets U, V ⊂ R (such that V × U is a neigborhood of (c, 0)) and a unique real analytic function K : U → V solution to equation (106). Furthermore this K is the unique solution to equation (106) of the form K : U → R since (c, 0) is the unique solution to equation (107) of the form (·, 0) ∈ R2. +∞ (n) n The fact that K(λ)= n=0 K λ is the unique solution to equation (106) in a neighborhood of 0 implies that P +∞ d K(1) = (λg(K(λ))) = g(K(0)) = g(k)(K(0))k dλ λ=0 k=0 X and also, for any n > 2, 1 dn K(n) = (λg(K(λ))) n! dλn λ=0 n−1 n d = g(K(λ)) n! dλn−1 λ=0 +∞

= g( k)K(ℓ1) ··· K(ℓk). (108) Xk=1 ℓ1+···+Xℓk=n−1 +∞ (n) n (n) (n) Suppose that fC P Gg,c(fC(·), ·), with fC(λ) = n=0 f λ , then we want to prove that |f | 6 K by induction on n. For n =0 we have P (0) (0) |f | 6 Gg,c(f(0), 0) = c = K .

For n =1 we have

+∞ +∞ (1) d (k) (0) k (k) (0) k (1) |f | 6 (G (fC(λ), λ)) = g(fC(0)) = g (f ) 6 g (K ) = K . dλ g,c kX=0 Xk=0

56 Supposing that |f (k)| 6 K(k) for any k 2, then we will have

n (n) 1 d |f | 6 G (fC(λ), λ) n! dλn g,c λ=0 n−1 n d 6 g(fC(λ)) n! dλn−1 λ=0 +∞

6 g( k)f (ℓ1) ··· f (ℓk) Xk=1 ℓ1+···+Xℓk=n−1 +∞ 6 g(k)K(ℓ1) ··· K(ℓk) Xk=1 ℓ1+···+Xℓk=n−1 6 K(n) where we used equality (108) and the induction hypothesis. This proves that fC P K. ✷

Remark 100 From formula (108) it is clear that the map (g,c) 7→ K is monotone, i.e. if K1(λ) =

Gg1,c1 (K1(λ), λ), K2(λ)= Gg2,c2 (K2(λ), λ) and g1 P g2 and c1 6 c2 then K1 P K2.

∂nΨ˜ s ε,λ2,t ˜ s We want to use Lemma 99 for giving an apriori bound on the norm of ∂λn (with Ψε,λ2,t as in ˜ s equation (102)) when λ = 0. This would imply that the Ψε,λ2,t is real analytic in λ when λ is small enough.

First of all we introduce the notion of real-analyticity for functions taking values in a Banach space.

Definition 101 Let X be a Banach space and let Z : U → X be an infinitly differentiable function (where U ⊂ R is an open set). We say that Z is real analytic in U if for any λ0 ∈ U we have

+∞ 1 dnZ Z(λ)= (λ )(λ − λ )n (109) n! dλn 0 0 Xk=0 for |λ − λ0| > 0 small enough (where the series on the right hand side of (109) is supposed to converge absolutely in X). If Z : U → X is differentiable an infinite number of times at 0 we can define the formal power series

+∞ Z˜(λ) := Z(n)λn (110) n=0 X where Z(n) ∈ X is given by 1 dnZ Z(n) := (0) ∈ X. n! dλn We can consider also the formal power series SZ (λ) defined as

+∞ (n) n SZ (λ) := kZ kλ . (111) n=0 X It is useful to recall an equivalent characterization of real analytic functions in Banach spaces.

Remark 102 We have that Z˜ (as given by (110)) is convergent in a neighborhood of 0 (and so it defines an analytic function in a neighborhood of 0) if and only if SZ (as given by (111)) is real analytic. Furthermore for any 0 P g, we have Z˜ P g if and only if SZ P g. It is important to note that if Z˜ is convergent (and so it defines an analytic function in a neighborhood of 0) this is not enough for proving that Z is also real analytic in a neighborhood of 0. Indeed, in general, if Z˜ is real analytic we could have Z˜(λ) 6= Z(λ) for λ 6=0. In order to overcome this problem we introduce the following useful theorems.

57 Theorem 103 Let X be a Banach space and let Z : U ⊂ R → X be a function defined in a neighborhood of 0. Then Z is real-analytic in U if and only if, for any ℓ∗ ∈ X∗, the function λ 7→ ℓ∗(Z(λ)) is real analytic in U. Proof The proof can be found in [38] Proposition 1 and Remark 2. ✷

Theorem 104 Suppose that Zn(λ) is a sequence of functions from an open neighborhood U ⊂ R of 0 into a Banach space X, converging pointwise Zn(λ) → Z(λ) to some function from U into X. Suppose furthermore that each Zn(λ) is real analytic and that there is a real analytic function 0 P K : U → R such that Zn P K, then Z(λ) is real analytic in a neighborhood of 0 and Z E K.

Proof ∗ ∗ ∗ ∗ 6 ∗ ∗ Consider ℓ ∈ X such that kℓ kX 1 and define ℓZn (λ) := ℓ (Zn(λ)). Since Zn are real ∗ ∗ ∗ ∗ analytic also ℓZn is real analytic, and also ℓZn (λ) → ℓZ (λ) := ℓ (Z(λ)) for any λ ∈ U. Furthermore ∗ P P ℓZn SZn K which implies that, by Remark 98, for any R such that K(R) is well defined, we have

∗ sup |ℓZn (λ)| 6 K(R), λ∈C,|λ|6R

∗ i.e. ℓZn are uniformly bounded in a (complex)-neighborhood of 0. This means by Vitali-Porter theorem ∗ ∗ (see, i.e. [146] Chapter 2, Section 2.4) that also ℓZ (λ) is real analytic for λ ∈ U ∩{|λ| < R}. Since ℓZ is real ∗ ∗ ∗ analytic for any ℓ ∈ X such that kℓ kX∗ =1, the function ℓ˜(Z(λ)) is real analytic in U ∩{|λ| < R} which implies, by Theorem 103, that Z is real analytic in U ∩ {|λ| < R} which is a non-empty neighborhood of 0. Finally we have

k k ∗ dkℓ∗ d SZ d ℓZ Zn (k) k (0) = sup k (0) 6 sup k (0) 6 K ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ N dλ ℓ ∈X ,kℓ kX =1 dλ ℓ ∈X ,kℓ kX =1,n∈ dλ

∗ ∗ dkℓ dkℓ Zn Z ✷ where we used that, from the Vitali-Porter theorem, dλk converges to dλk uniformly.

Remark 105 It is easy to generalize the proof of Theorem 104 replacing the convergence of Zn(λ) to Z(λ) ∈ X in the norm of X, with the convergence of Zn(λ) to Z(λ) ∈ X with respect to a (weaker) norm of a Banach space X′ ⊃ X in which X is densely contained.

k Let P : X → X be a polynomial map, i.e. there are some P0, P1,...,Pn such that Pk : X → X are continuous k-linear symmetric maps such that

n P (x) := Pk(x,...,x). Xk=0 We also suppose that there is P : R → R defined as

n k P(λ) := kPkkL(Xk,X)λ Xk=0 k where k·kL(Xk,X) is the natural operator norm of L(X ,X) that is the space of k-linear maps between k X into X. Suppose that ZP,C : U → X (where U ⊂ R is a neighborhood of 0) is the unique solution to the equation ZP,C (λ)= λP (Z(λ)) + C (112) where λ ∈ U and C ∈ X. We are now ready to prove a general apriori estimate for equations of the form (112).

Theorem 106 Suppose that a solution ZP,C to equation (112) is differentiable an infinite number of time in λ =0, then we have ZP,C P KP,C where KP,C is the unique analytic solution (in a neighborhood of 0) to the equation KP,C(λ)= λP(KP,C(λ)) + kCkX . (113)

58 Proof If we take the derivatives in the variable λ in the point λ =0 and write 1 dnZ Z(n) := P,C (0) ∈ X, P,C n! dλn we obtain from the formal series (110) and equation (112)

(0) (1) (0) ZP,C = C,ZP,C = P (Z )= P (C) and, for r > 2, n (r) (j1) (jk) ZP,C = Pk(ZP,C ,...,ZP,C ). j +···+j =r−1 kX=0 1 Xk This implies that (0) (1) kZP,CkX = kCkX , kZP,Ck 6 P(kCkX ) and, for r > 2,

n (r) (j1) (j2k+1) kZP,Ck 6 kP2k+1kL(X2k+1,X)kZP,C kX ···kZP,C kX . (114) j +···+j =r−1 kX=0 1 X2k+1

+∞ (n) n If we set SZP,C = n=0 kZP,Ckλ , then we have

P +∞ (k) r SZP,C (λ) = kCkX + kZP,Ckλ r=1 X +∞ n r (j1) (jk) P kCkX + λ kPkkL(Xk,X)kZP,C kX ···kZP,C kX r=1 j +···+j =r−1 X kX=0 1 Xk n +∞ r (j1) (jk) P kCkX + λ λ kPkkL(Xk,X)kZP,C kX ···kZP,C kX r=0 j +···+j =r Xk=0 X 1 Xk P kCkX + λP(SZP,C (λ)) where we used inequality (114) and the fact that, by Theorem 103, the function SP,C is real analytic in a neighborhood of 0 with positive derivatives in 0. By Lemma 99 the thesis follows by taking Gg,c(a, λ)= λP(a)+ kCkX . ✷ Proof of Theorem 93 We want to apply the previous theorems for proving Theorem 93. Here the Banach space X is X = C0(R, Cq(R2, A⊗ C4)) with the norm given by ˜ ˜ kΨkX = sup kΨkCq (R2,A⊗C4). t∈R

˜ ˜ The function Z ˜A (λ) is real analytic in λ in its radius of convergence. Furthermore it is a PN (Fε,hR,Y ),Bε,R,t solution to equation (104), which is of the form (112) with

t 2 ˜ (∆−mf )(t−τ) ˜ P3,hR,N (Ψ)(t, ·)= e PN (Fε,hR,Y (Ψ(τ, ·)))dτ, (115) Z−∞ A 3 ˜ ˜ 3 ˜ C = Bε,R,t, and PhR,N (λ)= kP3,hR,N kL(X ,X)λ . This means that, by Theorem 106, we have

Z ˜A P K ˜A , PN (Fε,hR,Y ),Bε,R,t P3,hR,N ,Bε,R where K ˜A is the solution to the equation P3,h,N ,Bε,R

3 A ˜ ˜ 3 ˜ ˜ K ˜A (λ)= λkP3,hR,N kL(X ,X) K ˜A (λ) + kBε,RkX . P3,hR,N ,Bε,R P3,hR,N ,Bε,RX   59 3 3 Since kP3,hR,N kL(X ,X) 6 kP3,hR kL(X ,X) (i.e. the polynomial P3,hR,N without the finite dimensional pro- jection), by Remark 100, we have Z ˜A P K ˜A . Since, by Theorem 65, Z ˜A PN (Fε,hR,Y ),Bε,R,t P3,hR ,Bε,R PN (Fε,hR,Y ),Bε,R,t converges pointwise to Z ˜A (i.e. the solution to equation (104) without the finite dimen- Fε,hR,Y ,Bε,R,t sional projection in the non-linearity) we have that, by Theorem 104, Z ˜A is real analytic and Fε,hR,Y ,Bε,R,t Z ˜A P K ˜A . Fε,hR,Y ,Bε,R,t P3,hR ,Bε,R ˜A ˜A ˜A Using again Remark 100 and the fact that kBε,R,tk 6 kBε,tkX (where, we recall, that Bε,R,t is the ˜A periodized version of Bε,t), we get Z ˜A P K ˜A P K ˜A . Furthermore, since, by Fε,hR,Y ,Bε,R,t P3,hR ,Bε,R P3,h,Bε Theorem 69, Z ˜A converges to Z ˜A (where Z ˜A is the solution to equation (104) Fε,hR,Y ,Bε,R,t Fε,h,Y ,Bε,t Fε,h,Y ,Bε,t ˜A ˜A ˜A ′ 0 q 2 4 where Bε,R,t is replaced by the non-periodized version Bε,t of Bε,R,t) in X = C (R, Cℓ (R , A⊗ C )) ⊃ X q (where we consider the weighted space Cℓ instead of the unweighted one), by Theorem 104 and Remark ˜ 105, we obtain that Z ˜A is real analytic with respect to λ and also Z ˜A P K ˜A . The Fε,h,Y ,Bε,t Fε,h,Y ,Bε,t P3,h,Bε thesis follows in a similar way, using the fact that, by Remark 100 and the fact that kP3,hkL(X3,X) 6 kP3kL(X3,X) (where P3 is the nonlinearity defined in equation (115) where all the cut-offs have been removed), we have K ˜A P K ˜A , and the convergence Z ˜A to Z ˜A (i.e. the solution P3,h,Bε P3,Bε Fε,h,Y ,Bε,t Fε,Y ,Bε,t to equation (102)). ✷

E Other interacting models

In this last appendix we want to give an idea of other Euclidean quantum fields models which can be approached within the theory developed in this paper.

Non-local, Yukawa2-like models. The discussion in Section 5 can be extended in a straightforward way to other fermionic quantum field theories. One of them is two dimensional quantum electrodynamics 2 4 (QED2), see, e.g. [114] Chapter 3. In this case we consider a fermionic (free) field ψ˜ ∈ S(R , A⊗ C ) with the same covariance as in (63) and (64), and two (massive-free-scalar) bosons A = (A1, A2) ∈ S(R2, A⊗ R2) independent of each other with covariance

i j 2 −1 hA (x)A (y)i = δi,j (−∆+ M ) (x − y), where i, j ∈{1, 2}, and M ∈ R+ is their mass. The interaction potential is given (formally) by

i VQED(A, ψ˜)= λ (ψ¯(x) · γiψ(x))A (x)dx, R2 i=1,2 X Z where λ ∈ R, which, when the bosons A are integrated out, gives rise to the effective potential

2 V(ψ˜)= λ (ψ¯(x) · γiψ(x))GM (x − y)(ψ¯(y) · γiψ(y))dxdy (116) R4 i=1,2 X Z 2 −1 where GM (x − y) = (−∆+ M ) (x − y). If we consider a regularized version ψ˜ε(x) = ψ˜(aε(·− x)), ε > 0, as it was done in Section 5, of the fermion field ψ˜, with the purely fermionic interaction V, the stochastic quantization equation is of the form (70), where the non-linearity Fε,Y = Fε,1,Y is replaced by the following term

a∗2 ∗ −Ψ¯ γ G (y − ·)Ψ¯ (y) · γ Ψ (y)dy F (Ψ˜ )= C i=1,2 ε ε,t i R2 M ε,t i ε,t , ε,QED ε,t f a∗2 ∗ γ Ψ G (y − ·)Ψ¯ (y) · γ Ψ (y)dy  Pi=1,2 ε i ε,t RR2 M ε,t i ε,t   P R  and where Cf is defined in eq. (71). It is simple to generalize Lemma 61 for the nonlinearity Fε,QED and this permits to prove an analogous version of Theorem 59 using the same methods as in Section 5. An analogous reasoning can be extended also to the Yukawa model with pseudo-scalar current having an effective potential of the form

2 V(ψ˜)= λ (ψ¯(x) · γ5ψ(x))G(x − y)(ψ¯(y) · γ5ψ(y))dxdy, R4 Z

60 2 −1 where γ5 = iγ1γ2 and G(x − y) = (−∆+ mb ) (x − y) as in eq. (68).

Local models. The treatment proposed for the regularized Yukawa model can extended also to purely fermionic models with quartic interaction. We describe as an example the Gross–Neveu model (see [114] Chapter 6 for this model) in two dimensions. In this case we consider a (free) fermionic Dirac field ψ˜ with interaction given by the potential

V(ψ˜)= −λ2 (ψ¯(x) · ψ(x))2dx. R2 Z The main difference in the corresponding stochastic quantization equation, compared with that of the QED2 model above, is that Fε,QED is replaced by

a∗2 ∗ (−Ψ¯ (Ψ¯ · Ψ )) F (Ψ˜ )= C ε ε,t ε,t ε,t ε,GN ε,t f a∗2 ∗ (Ψ (Ψ¯ · Ψ ))  ε ε,t ε,t ε,t  where Cf is defined, as before, in eq. (71). Also for Fε,GN a generalization of Lemma 61 holds. This allows us to prove an analogous version of Theorem 59 using the same methods as in Section 5.

Supersymmetric models. Supersymmetric quantum field models describe quantum fields where the bosonic and fermionic fields are related, one to the other, by a special transformation called supersymme- try. It is not possible to reduce a supersymmetric non free (i.e. interacting) model to a purely fermionic one since the presence of interaction and supersymmetry forces the potential to have a part of bosonic self interaction. Nevertheless from an analytical point of view it can be interesting to study supersymmetric models considering free bosons and integrating them out as was done in Section 5. One of the most important examples of super-symmetric models is Wess–Zumino model composed by two scalar bosons interacting with a Dirac fermionic field through a Yukawa scalar and pseudo-scalar types interaction (see [90, 12, 80]). This model can also be handled as in Section 5. Another source of supersymmetric quantum field models is the supersymmetric formulation of stochastic differential equations (see [180] Chapter 15). In particular here we discuss the model related with elliptic stochastic differential equations (see [101, 100, 4] see also [48]). In this case we have a fermionic field ψ˜ = (ψ, ψ¯) ∈ S(R2, A⊗ R2) with covariance

hψ¯(x)ψ(y)i = (−∆+ m2)−1(x − y), where m ∈ R+ is a positive constant. The corresponding effective potential V, obtained after integrating over the boson field, is

2 V(ψ˜)= λ (ψ¯(x)ψ(x))Km(x − y)(ψ¯(y)ψ(y))dxdy, R4 Z 2 −2 where Km(x − y) := (−∆+ m ) (x − y). The stochastic quantization equation reads

2 2 dΨ˜ ε,t = ((−∆+ m )Ψ˜ ε,t + λ Fε,E(Ψ˜ ε,t))dt + dBˆε,t (117) where Bˆε,t = aε∗(Bt, B¯t), with aε as in Section 5.1, (Bt, B¯t) is a couple of Brownian motion, i.e. they 0 1 are the integral of a Gaussian noise with covariance J = , and F is defined by −1 0 ε,E   a∗2 ∗ −Ψ¯ K (y − ·)Ψ¯ (y)Ψ (y)dy F (Ψ˜ )= −J · ε ε,t R2 m ε,t ε,t . (118) ε,E ε,t a∗2 ∗ Ψ K (y − ·)Ψ¯ (y)Ψ (y)dy  ε ε,t RR2 m ε,t ε,t   Also in this case a generalization of Lemma 61 andR of Theorem 59 can be proven.

Comments. As it was shown, the theory presented in this article is able to construct the regularized version of many models of fermionic quantum field theories. The mathematical differences between these models can only be seen when the regularization is removed. More precisely the solution to the supersymmetric model (117) and (118) is expected to be a distribution of regularity C−δ(R2, A⊗ R2) (for 4 any δ > 0) as in Φ2 bosonic case. The non-local interactions (resulting after the bosons are integrated

61 out) such as in the Yukawa model (in Section 5) and in the QED model (defined by eqs. (116)) have solutions in C−1/2−δ(R2, A⊗ R2) (for any δ > 0) and they are (formally) super-renormalizable much like 4 the Φ3 bosonic model. The main difference between Yukawa and QED being that in the Yukawa case the mass mb of the boson ϕ used for defining G goes to +∞ as the regularisation is removed and have to be renormalized, while in the QED setting the mass of the photon M can be kept constant. Finally, let us also mention that the removal of the regularization ε in the study of the Gross–Neveu model in our setting would be particularly interesting (the model being only renormalizable and not superrenormalizable as are the other ones, it remains to see whether it is “asymptotic free” character, according to physical terminology, will help this construction).

References

[1] Luigi Accardi. An outline of quantum probability. Preprint 10.13140/rg.2.1.3078.3844, 2015.

[2] Luigi Accardi, Alberto Frigerio, and John T. Lewis. Quantum stochastic processes. Ky- oto University. Research Institute for Mathematical Sciences. Publications, 18(1):97–133, 1982. 10.2977/prims/1195184017. [3] , Francesco C. De Vecchi, and Massimiliano Gubinelli. The elliptic stochastic quantization of some two dimensional Euclidean QFTs. ArXiv:1906.11187 [math-ph], jun 2019. ArXiv: 1906.11187, to appear in Annales de l’Institut Henri Poincaré, Probabilités et Statistiques. [4] Sergio Albeverio, Francesco C. De Vecchi, and Massimiliano Gubinelli. Elliptic stochastic quanti- zation. Ann. Probab., 48(4):1693–1741, 2020.

4 [5] Sergio Albeverio and Seiichiro Kusuoka. The invariant measure and the flow associated to the Φ3- quantum field model. Annali della Scuola Normale di Pisa - Classe di Scienze, 2018. 10.2422/2036- 2145.201809_008.

[6] Daniel Alpay, Ismael L. Paiva, and Daniele C. Struppa. Distribution spaces and a new construction of stochastic processes associated with the Grassmann algebra. Journal of Mathematical Physics, 60(1):13508–21, 2019. 10.1063/1.5052010. [7] Herbert Amann. Operator-valued Fourier multipliers, vector-valued Besov spaces, and applications. Math. Nachr., 186:5–56, 1997. 10.1002/mana.3211860102. [8] Herbert Amann. Linear and quasilinear parabolic problems. Vol. II, volume 106 of Monographs in Mathematics. Birkhäuser/Springer, Cham, 2019. 10.1007/978-3-030-11763-4. [9] G. F. De Angelis, G. Jona-Lasinio, and V. Sidoravicius. Berezin integrals and Poisson pro- cesses. Journal of Physics A: Mathematical and General, 31(1):289–308, jan 1998. 10.1088/0305- 4470/31/1/026.

[10] D. B. Applebaum and R. L. Hudson. Fermion Itô’s formula and stochastic evolutions. Comm. Math. Phys., 96(4):473–496, 1984. [11] D. Applebaum and R. L. Hudson. Fermion diffusions. J. Math. Phys., 25(4):858–861, 1984. 10.1063/1.526236. [12] Asao Arai. A general class of infinite-dimensional Dirac operators and path integral representation of their index. J. Funct. Anal., 105(2):342–408, 1992. [13] Huzihiro Araki. On the Diagonalization of a Bilinear Hamiltonian by a Bogoliubov Transforma- tion. Publications of the Research Institute for Mathematical Sciences, Kyoto University. Ser. A, 4(2):387–412, 1968. 10.2977/prims/1195194882.

[14] Huzihiro Araki. On Quasifree States of CAR and Bogoliubov Automorphisms. Publications of the Research Institute for Mathematical Sciences, 6(3):385–442, 1970. 10.2977/prims/1195193913.

62 [15] Huzihiro Araki. Mathematical theory of quantum fields, volume 101 of International Series of Monographs on Physics. Oxford University Press, New York, 1999. Translated from the 1993 Japanese original by Ursula Carow-Watamura.

[16] Huzihiro Araki and Walter Wyss. Representations of canonical anticommutation relations. Helvetica Physica Acta, 37(2):136, 1964.

[17] John C. Baez, Irving E. Segal, and Zheng-Fang Zhou. Introduction to algebraic and constructive quantum field theory. Princeton Series in Physics. Princeton University Press, Princeton, NJ, 1992. 10.1515/9781400862504. [18] Dorothea Bahns, Sergio Doplicher, Gerardo Morsella, and Gherardo Piacitelli. Quantum spacetime and algebraic quantum field theory. In Advances in algebraic quantum field theory, Math. Phys. Stud., pages 289–329. Springer, Cham, 2015. 10.1007/978-3-319-21353-87.

[19] C. Barnett, R. F. Streater, and I. F. Wilde. The Itô-Clifford integral. J. , 48(2):172–212, 1982. 10.1016/0022-1236(82)90066-0. [20] C. Barnett, R. F. Streater, and I. F. Wilde. The Itô-Clifford integral. II. Stochastic differential equations. J. London Math. Soc. (2), 27(2):373–384, 1983. 10.1112/jlms/s2-27.2.373. [21] C. Barnett, R. F. Streater, and I. F. Wilde. The Itô-Clifford integral. III. The Markov property of solutions to stochastic differential equations. Comm. Math. Phys., 89(1):13–17, 1983.

[22] Guy A. Battle and Lon Rosen. On the infinite volume limit of the strongly coupled Yukawa2 model. Journal of Mathematical Physics, 22(4):770–776, apr 1981. 10.1063/1.524982. [23] Hellmut Baumgärtel. Operator algebraic methods in quantum field theory. Akademie Verlag, Berlin, 1995. A series of lectures.

[24] Hellmut Baumgärtel and Manfred Wollenberg. Causal nets of operator algebras, volume 80 of Mathematische Lehrbücher und Monographien, II. Abteilung: Mathematische Monographien [Math- ematical Textbooks and Monographs, Part II: Mathematical Monographs]. Akademie-Verlag, Berlin, 1992. Mathematical aspects of algebraic quantum field theory.

[25] Viacheslav P. Belavkin. Quantum stochastic calculus and quantum nonlinear filtering. J. Multi- variate Anal., 42(2):171–201, 1992. 10.1016/0047-259X(92)90042-E. [26] G. Benfatto, P. Falco, and V. Mastropietro. Functional Integral Construction of the Massive Thirring model: Verification of Axioms and Massless Limit. Communications in Mathematical Physics, 273(1):67–118, jul 2007. 10.1007/s00220-007-0254-y. [27] Y. M. Berezansky and Y. G. Kondratiev. Spectral methods in infinite-dimensional analysis. Vol. 1, volume 12/1 of Mathematical Physics and . Kluwer Academic Publishers, Dordrecht, 1995. Translated from the 1988 Russian original by P. V. Malyshev and D. V. Malyshev and revised by the authors. 10.1007/978-94-011-0509-5.

[28] F. A. Berezin. The method of second quantization. Translated from the Russian by Nobumichi Mugibayashi and Alan Jeffrey. Pure and Applied Physics, Vol. 24. Academic Press, New York- London, 1966.

[29] Felix Alexandrovich Berezin. Introduction to superanalysis, volume 9 of Mathematical Physics and Applied Mathematics. D. Reidel Publishing Co., Dordrecht, 1987. 10.1007/978-94-017-1963-6. [30] Lorenzo Bertini, Giovanni Jona-Lasinio, and Claudio Parrinello. Stochastic Quantization, Stochas- tic Calculus and Path Integrals: Selected Topics. Progress of Theoretical Physics Supplement, 111:83–113, jan 1993. 10.1143/PTPS.111.83.

[31] Philippe Biane. Calcul stochastique non-commutatif. In Pierre Bernard, editor, Lectures on Prob- ability Theory, volume 1608, pages 1–96. Springer Berlin Heidelberg, Berlin, Heidelberg, 1995. 10.1007/BFb0095746.

63 [32] N. N. Bogolubov, Anatoly A. Logunov, A. I. Oksak, and I. Todorov. General Principles of Quantum Theory. Mathematical Physics and Applied Mathematics. Springer Netherlands, 1990. [33] Frank F. Bonsall and John Duncan. Complete normed algebras. Springer-Verlag, New York- Heidelberg, 1973.

[34] Luigi Marcello Borasi. Probabilistic and differential geometric methods for relativistic and Euclidean Dirac and radiation fields. PhD thesis, University of Bonn, Bonn, Germany, jul 2019. [35] Marek Bożejko and Roland Speicher. Interpolations between bosonic and fermionic relations given by generalized brownian motions. Mathematische Zeitschrift, 222(1):135–160, may 1996. 10.1007/BF02621861.

[36] Ola Bratteli and Derek W. Robinson. Operator algebras and quantum statistical mechanics. 1. Texts and Monographs in Physics. Springer-Verlag, New York, Second edition, 1987. 10.1007/978- 3-662-02520-8.

[37] Ola Bratteli and Derek W. Robinson. Operator algebras and quantum statistical mechanics. 2. Texts and Monographs in Physics. Springer-Verlag, Berlin, Second edition, 1997. 10.1007/978-3- 662-03444-6.

[38] Felix E. Browder. Analyticity and partial differential equations. I. Amer. J. Math., 84:666–710, 1962.

[39] Eduardo R. Caianiello. and Renormalization in , volume 38. Benjamin, 1973.

[40] Eric Carlen and Paul Krée. Lp estimates on iterated stochastic integrals. Ann. Probab., 19(1):354– 368, 1991. [41] Rémi Catellier and Khalil Chouk. Paracontrolled distributions and the 3-dimensional stochastic quantization equation. The Annals of Probability, 46(5):2621–2679, 2018. 10.1214/17-AOP1235. [42] Fabio Cipriani. Noncommutative : a survey. J. Geom. Phys., 105:25–59, 2016. 10.1016/j.geomphys.2016.03.016.

[43] Alain Connes. Noncommutative geometry. Academic Press, San Diego, 1994. [44] Alan Cooper and Lon Rosen. The Weakly Coupled Yukawa$_2$ Field Theory: Cluster Expansion and Wightman Axioms. Transactions of the American Mathematical Society, 234(1):1–88, 1977. 10.2307/1997994.

[45] Leo Corry. Modern Algebra and the Rise of Mathematical Structures. Birkhäuser Basel, 2 edition, 2004. 10.1007/978-3-0348-7917-0.

[46] P.H. Damgaard and K. Tsokos. Stochastic quantization with fermions. Nuclear Physics B, 235(1):75–92, may 1984. 10.1016/0550-3213(84)90149-4.

[47] Poul Henrik Damgaard and Helmuth Hüffel. Stochastic Quantization. World Scientific, 1988. [48] Francesco C. De Vecchi and Massimiliano Gubinelli. A note on supersymmetry and stochastic differential equations. ArXiv:1912.04830 [math-ph], dec 2019. ArXiv: 1912.04830. [49] Bryce DeWitt. Supermanifolds. Cambridge Monographs on Mathematical Physics. Cambridge University Press, Cambridge, Second edition, 1992. 10.1017/CBO9780511564000.

[50] Jonathan Dimock. Estimates, renormalized currents, and field equations for the Yukawa2 field theory. Annals of Physics, 72(1):177–242, jul 1972. 10.1016/0003-4916(72)90241-2. [51] Sjoerd Dirksen. Noncommutative stochastic integration through decoupling. Journal of Mathemat- ical Analysis and Applications, 370(1):200–223, oct 2010. 10.1016/j.jmaa.2010.04.062.

64 [52] M. Disertori and V. Rivasseau. Continuous Constructive Fermionic Renormalization. Annales Henri Poincaré, 1(1):1–57, feb 2000. 10.1007/PL00000998. [53] A. N. Efremov. Stochastic Quantization of Massive Fermions. International Journal of Theoretical Physics, 58(4):1150–1156, apr 2019. 10.1007/s10773-019-04006-w. [54] Gérard G. Emch. Mathematical and conceptual foundations of 20th-century physics, volume 100 of North-Holland Mathematics Studies. North-Holland Publishing Co., Amsterdam, 1984. Notas de Matemática [Mathematical Notes], 100. [55] Paul Federbush and Basilis Gidas. Renormalization of the one-space dimensional Yukawa model by unitary transformation. Ann. Physics, 68:98–101, 1971. [56] Joel Feldman, Horst Knörrer, and Eugene Trubowitz. Fermionic functional integrals and the renor- malization group, volume 16 of CRM Monograph Series. American Mathematical Society, Provi- dence, RI, 2002. [57] Joel Feldman, J. Magnen, V. Rivasseau, and R. Sénéor. A renormalizable field theory: The massive Gross-Neveu model in two dimensions. Commun.Math. Phys., 103(1):67–103, mar 1986. 10.1007/BF01464282.

[58] . Mathematical Aspects of the Quantum Theory of Fields. Interscience Pub- lishers, 1953.

[59] Jürg Fröhlich and Konrad Osterwalder. Is There a Euclidean Field Theory for Fermions. Helv. Phys. Acta, 47:781, 1975. [60] T. Fukai, H. Nakazato, I. Ohba, K. Okano, and Y. Yamanaka. Stochastic Quantization Method of Fermion Fields. Progress of Theoretical Physics, 69(5):1600–1616, may 1983. 10.1143/PTP.69.1600. [61] K. Gawędzki and A. Kupiainen. Gross-Neveu model through convergent perturbation expansions. Communications in Mathematical Physics, 102(1):1–30, 1985. [62] I. Gelfand and M. Neumark. On the imbedding of normed rings into the ring of operators in Hilbert space. Rec. Math. [Mat. Sbornik] N.S., 12(54):197–213, 1943.

[63] James Glimm and Arthur Jaffe. The Yukawa2 quantum field theory without cutoffs. J. Functional Analysis, 7:323–357, 1971. [64] James Glimm and Arthur Jaffe. Quantum Field Theory Models. In James Glimm and Arthur Jaffe, editors, Quantum Field Theory and Statistical Mechanics: Expositions, pages 11–121. Birkhäuser, Boston, MA, 1985. 10.1007/978-1-4612-5158-34.

[65] James Glimm and Arthur Jaffe. Quantum physics. Springer-Verlag, New York, Second edition, 1987. A functional integral point of view.

[66] Maria Gordina. Stochastic differential equations on noncommutative L2. In Finite and infinite dimensional analysis in honor of Leonard Gross (New Orleans, LA, 2001), volume 317 of Contemp. Math., pages 87–98. Amer. Math. Soc., Providence, RI, 2003. 10.1090/conm/317/05521. [67] Leonard Gross. Abstract Wiener spaces. In Proc. Fifth Berkeley Sympos. Math. Statist. and Probability (Berkeley, Calif., 1965/66), Vol. II: Contributions to Probability Theory, Part 1, pages 31–42. Univ. California Press, Berkeley, Calif., 1967.

[68] Leonard Gross. Existence and uniqueness of physical ground states. Journal of Functional Analysis, 10(1):52–109, may 1972. 10.1016/0022-1236(72)90057-2. [69] Leonard Gross. Hypercontractivity and logarithmic Sobolev inequalities for the Clifford-Dirichlet form. Duke Mathematical Journal, 42(3):383–396, sep 1975. 10.1215/S0012-7094-75-04237-4. [70] Leonard Gross. Euclidean fermion fields. In Les méthodes mathématiques de la théorie quantique des champs (Colloq. Internat. CNRS, No. 248, Marseille, 1975), pages 131–146. 1976.

65 [71] Leonard Gross. On the formula of Mathews and Salam. Journal of Functional Analysis, 25(2):162– 209, jun 1977. 10.1016/0022-1236(77)90039-8.

[72] Lonard Gross. A noncommutative extension of the Perron-Frobenius theorem. Bulletin of the American Mathematical Society, 77(3):343–348, may 1971. 10.1090/S0002-9904-1971-12686-1.

4 [73] Massimiliano Gubinelli and Martina Hofmanová. A PDE construction of the Euclidean Φ3 quantum field theory. ArXiv:1810.01700 [math-ph], oct 2018. ArXiv: 1810.01700. [74] Massimiliano Gubinelli and Martina Hofmanová. Global Solutions to Elliptic and Parabolic Φ4 Models in Euclidean Space. Communications in Mathematical Physics, 368(3):1201–1266, 2019. [75] Massimiliano Gubinelli, Peter Imkeller, and Nicolas Perkowski. Paracontrolled distributions and singular PDEs. Forum of Mathematics. Pi, 3:0, 2015. 10.1017/fmp.2015.2. [76] Massimiliano Gubinelli, Herbert Koch, and Tadahiro Oh. Paracontrolled approach to the three- dimensional stochastic nonlinear wave equation with quadratic nonlinearity. ArXiv:1811.07808 [math], nov 2018. ArXiv: 1811.07808. [77] Massimiliano Gubinelli, Herbert Koch, and Tadahiro Oh. Renormalization of the two-dimensional stochastic nonlinear wave equations. Transactions of the American Mathematical Society, page 1, 2018. 10.1090/tran/7452.

[78] Francesco Guerra. Local algebras in Euclidean quantum field theory. In Symposia Mathematica, Vol. XX (Convegno sulle Algebre C* e loro Applicazioni in Fisica Teorica, Convegno sulla Teoria degli Operatori Indice e Teoria K, INDAM, Roma, 1974), pages 13–26. 1976. [79] Rudolf Haag. Local quantum physics. Texts and Monographs in Physics. Springer-Verlag, Berlin, Second edition, 1996. 10.1007/978-3-642-61458-3.

[80] Z. Haba and J. Kupsch. Supersymmetry in Euclidean Quantum Field Theory. Fortschritte der Physik/Progress of Physics, 43(1):41–66, 1995. 10.1002/prop.2190430103. [81] M. Hairer. A theory of regularity structures. Inventiones mathematicae, 198(2):269–504, 2014. 10.1007/s00222-014-0505-4.

[82] K. Hepp. Théorie de la renormalisation. Cours donné à l’École Polytechnique, Paris. Lecture Notes in Physics, Vol. 2. Springer-Verlag, Berlin-New York, 1969.

[83] Takeyuki Hida, Hui-Hsiung Kuo, Jürgen Potthoff, and Walter Streit. White Noise: An Infinite Dimensional Calculus. Springer Science & Business Media, jun 2013. [84] Alexander Holevo. Probabilistic and statistical aspects of quantum theory, volume 1 of Quaderni/Monographs. Edizioni della Normale, Pisa, Second edition, 2011. 10.1007/978-88-7642- 378-9.

[85] Elton P. Hsu. Stochastic Analysis on Manifolds. Graduate Studies in Mathematics 38. American Mathematical Society, 2002. [86] R. L. Hudson and K. R. Parthasarathy. Unification of fermion and Boson stochastic calculus. Communications In Mathematical Physics, 104(3):457–470, sep 1986. 10.1007/BF01210951. [87] I. A. Ignatyuk, V. A. Malyshev, and V. Sidoravičius. Convergence of a Method of the Stochastic Quantization II. Theory of Probability & Its Applications, 37(4):599–620, jan 1993. 10.1137/1137117.

[88] Claude Itzykson and Jean Bernard Zuber. Quantum field theory. McGraw-Hill International Book Co., New York, 1980. International Series in Pure and Applied Physics.

[89] V. D. Ivashchuk. Infinite-dimensional Grassmann-Banach algebras. ArXiv:math-ph/0009006, sep 2000. ArXiv: math-ph/0009006.

66 [90] Arthur Jaffe, Andrzej Lesniewski, and Jonathan Weitsman. The two-dimensional, N = 2 Wess- Zumino model on a cylinder. Comm. Math. Phys., 114(1):147–165, 1988. [91] Svante Janson. Gaussian Hilbert Spaces. Cambridge University Press, jun 1997. [92] G. Jona-Lasinio. Stochastic quantization: a new domain for stochastic analysis. In Proceedings of the 1st World Congress of the Bernoulli Society, Vol. 1 (Tashkent, 1986), pages 535–546. VNU Sci. Press, Utrecht, 1987.

[93] G. Jona-Lasinio and P. K. Mitter. On the stochastic quantization of field theory. Communications in Mathematical Physics (1965-1997), 101(3):409–436, 1985.

4 [94] G. Jona-Lasinio and P. K. Mitter. Large deviation estimates in the stochastic quantization of φ2. Communications in Mathematical Physics, 130(1):111–121, may 1990. 10.1007/BF02099877. [95] Res Jost. The general theory of quantized fields, volume 1960 of Mark Kac, editor. Lectures in Applied Mathematics (Proceedings of the Summer Seminar, Boulder, Colorado. American Mathe- matical Society, Providence, R.I., 1965.

[96] Richard V. Kadison and John R. Ringrose. Fundamentals of the theory of operator algebras. Vol. I, volume 15 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 1997. Elementary theory, Reprint of the 1983 original.

[97] Richard V. Kadison and John R. Ringrose. Fundamentals of the theory of operator algebras. Vol. II, volume 16 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 1997. Advanced theory, Corrected reprint of the 1986 original. 10.1090/gsm/016/01. [98] Yujiro Kakudo, Yukio Taguchi, Azuma Tanaka, and Kunio Yamamoto. Gauge-independent calcu- lation of S-matrix elements in quantum electrodynamics. Progress of Theoretical Physics (Kyoto), 69(4):1225–1235, 1983.

[99] A. Yu Khrennikov. Equations on a superspace. Mathematics of the USSR-Izvestiya, 36(3):597, 1991. Publisher: IOP Publishing. 10.1070/IM1991v036n03ABEH002036.

[100] Abel Klein. Supersymmetry and a two-dimensional reduction in random phenomena. In Quantum probability and applications, II (Heidelberg, 1984), volume 1136 of Lecture Notes in Math., pages 306–317. Springer, Berlin, 1985. 10.1007/BFb0074481. [101] Abel Klein and J. Fernando Perez. Supersymmetry and dimensional reduction: A non-perturbative proof. Physics Letters B, 125(6):473–475, jun 1983. 10.1016/0370-2693(83)91329-1. [102] Burkhard Kümmerer. Survey on a theory of noncommutative stationary Markov processes. In Quantum probability and applications, III (Oberwolfach, 1987), volume 1303 of Lecture Notes in Math., pages 154–182. Springer, Berlin, 1988. 10.1007/BFb0078061. [103] A. Kupiainen. and Stochastic PDEs. Annales Henri Poincaré, 17(3):497– 535, 2016. 10.1007/s00023-015-0408-y.

[104] J. Kupsch. Fermionic and supersymmetric stochastic processes. Journal of Geometry and Physics, 11(1):507–516, jun 1993. 10.1016/0393-0440(93)90074-O.

[105] O. E. Lanford III. Construction of quantum fields interacting by a cutoff Yukawa coupling. PhD thesis, Princeton university, jan 1966.

[106] Y. Le Jan. Temps local et superchamp. In Séminaire de Probabilités, XXI, volume 1247 of Lecture Notes in Math., pages 176–190. Springer, Berlin, 1987. 10.1007/BFb0077633. [107] Detlef Lehmann. A probabilistic approach to Euclidean Dirac fields. Journal of Mathematical Physics, 32(8):2158–2166, aug 1991. Publisher: American Institute of Physics. 10.1063/1.529189. [108] D. A. Leites. Introduction to the theory of supermanifolds. Uspekhi Mat. Nauk, 35(1(211)):3–57, 1980.

67 [109] Steven Leppard and Alice Rogers. A Feynman-Kac formula for anticommuting Brownian mo- tion. Journal of Physics A: Mathematical and General, 34(3):555–568, jan 2001. 10.1088/0305- 4470/34/3/315.

[110] A. Lesniewski. Effective action for the Yukawa2 quantum field theory. Communications in Mathe- matical Physics, 108(3):437–467, sep 1987. 10.1007/BF01212319.

[111] J. Magnen and R. Sénéor. The Wightman axioms for the weakly coupled Yukawa model in two dimensions. Comm. Math. Phys., 51(3):297–313, 1976.

[112] J. Magnen and R. Sénéor. Yukawa quantum field theory in three dimensions (Y3). In Third International Conference on Collective Phenomena (Moscow, 1978), volume 337 of Ann. New York Acad. Sci., pages 13–43. New York Acad. Sci., New York, 1980.

[113] A. I. Markushevich. Theory of functions of a complex variable. Vol. II. Revised English edition translated and edited by Richard A. Silverman. Prentice-Hall, Inc., Englewood Cliffs, N.J., 1965.

[114] Vieri Mastropietro. Non-Perturbative Renormalization. World Scientific Publishing Co Pte Ltd, Hackensack, NJ, 2008.

[115] P. T. Matthews and A. Salam. Propagators of quantized field. Il Nuovo Cimento (1955-1965), 2(1):120–134, jul 1955. 10.1007/BF02856011.

[116] Oliver A. McBryan. Volume dependence of Schwinger function in the Yukawa2 quantum field theory. Communications in Mathematical Physics, 45(3):279–294, 1975. Publisher: Springer-Verlag.

[117] Oliver A. McBryan and Yong Moon Park. Lorentz covariance of the Yukawa2 quantum field theory. J. Mathematical Phys., 16:104–110, 1975. [118] Paul-André Meyer. Quantum probability for probabilists, volume 1538 of Lecture Notes in Mathe- matics. Springer-Verlag, Berlin, 1993. 10.1007/978-3-662-21558-6.

4 [119] A. Moinat and H. Weber. Space-time localisation for the dynamic Φ3 model. ArXiv:1811.05764, nov 2018. ArXiv: 1811.05764.

[120] J.-C. Mourrat and H. Weber. Global well-posedness of the dynamic Φ4 model in the plane. The Annals of Probability, 45(4):2398–2476, 2017. 10.1214/16-AOP1116.

4 [121] J.-C. Mourrat and H. Weber. The dynamic Φ3 model comes down from infinity. Comm. Math. Phys., 356(3):673–753, 2017. 10.1007/s00220-017-2997-4.

[122] M. A. Naimark. Normed algebras. Wolters-Noordhoff Publishing, Groningen, Third edition, 1972. Translated from the second Russian edition by Leo F. Boron, Wolters-Noordhoff Series of Mono- graphs and Textbooks on Pure and Applied Mathematics.

[123] Tadao Nakano. Quantum Field Theory in Terms of Euclidean Parameters. Progress of Theoretical Physics, 21(2):241–259, feb 1959. Publisher: Oxford Academic. 10.1143/PTP.21.241.

[124] Edward Nelson. Construction of quantum fields from Markoff fields. Journal of Functional Analysis, 12(1):97–112, jan 1973. 10.1016/0022-1236(73)90091-8.

[125] Edward Nelson. Notes on non-commutative integration. Journal of Functional Analysis, 15(2):103– 116, feb 1974. 10.1016/0022-1236(74)90014-7.

4 [126] Edward P. Osipov. Quantum interaction φ4, the construction of quantum field defined as a bilinear form. Journal of Mathematical Physics, 41(2):759–786, 2000. 10.1063/1.533162.

[127] Konrad Osterwalder. Euclidean fermi fields. In Constructive Quantum Field Theory, Lecture Notes in Physics, pages 326–331. Springer, Berlin, Heidelberg, 1973. 10.1007/BFb0113094. [128] Konrad Osterwalder and Robert Schrader. Euclidean Fermi fields and a Feynman-Kac formula for Boson-Fermions models. Helvetica Physica Acta, 46:277–302, 1973.

68 [129] B. G. Pachpatte. Inequalities for differential and integral equations, volume 197 of Mathematics in Science and Engineering. Academic Press, Inc., San Diego, CA, 1998. [130] John Palmer. Euclidean Fermi fields. Journal of Functional Analysis, 36(3):287–312, may 1980. 10.1016/0022-1236(80)90092-0.

[131] G. Parisi and Yong Shi Wu. without gauge fixing. Scientia Sinica. Zhongguo Kexue, 24(4):483–496, 1981. [132] K. R. Parthasarathy. An Introduction to Quantum Stochastic Calculus. Springer Science & Business Media, 1992.

[133] Vladimir G. Pestov. Analysis on superspace: an overview. Bulletin of the Australian Mathematical Society, 50(1):135–165, aug 1994. Publisher: Cambridge University Press. 10.1017/S0004972700009643.

[134] Roger Plymen and Paul Robinson. Spinors in Hilbert Space. Cambridge University Press, 1994. [135] Agostino Prastaro and Themistocles M. Rassias. Geometry in Partial Differential Equations. World Scientific, 1994.

[136] V. Rivasseau. From Perturbative to Constructive Renormalization. Princeton University Press, Princeton, N.J, 2 edition edition, may 1991.

[137] A. Rogers. Stochastic calculus in superspace. I. Supersymmetric Hamiltonians. Journal of Physics A: Mathematical and General, 25(2):447–468, jan 1992. 10.1088/0305-4470/25/2/024. [138] A. Rogers. Stochastic calculus in superspace. II. Differential forms, supermanifolds and the Atiyah- Singer index theorem. Journal of Physics A: Mathematical and General, 25(22):6043–6062, nov 1992. 10.1088/0305-4470/25/22/027. [139] Alice Rogers. Graded manifolds, supermanifolds and infinite-dimensional Grassmann algebras. Communications in Mathematical Physics, 105(3):375–384, sep 1986. 10.1007/BF01205932. [140] Alice Rogers. Path integration, anticommuting variables, and supersymmetry. Journal of Mathe- matical Physics, 36(5):2531–2545, may 1995. 10.1063/1.531049. [141] Alice Rogers. Supermanifolds. World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, 2007. 10.1142/9789812708854.

[142] Jonathan Rosenberg. Noncommutative variations on Laplace’s equation. Analysis & PDE, 1(1):95– 114, oct 2008. Publisher: Mathematical Sciences Publishers. 10.2140/apde.2008.1.95.

[143] Raymond A. Ryan. Introduction to tensor products of Banach spaces. Springer Monographs in Mathematics. Springer-Verlag London, Ltd., London, 2002.

[144] Manfred Salmhofer. Renormalization: An Introduction. Springer, Berlin ; New York, 1st Corrected ed. 1999, Corr. 2nd printing 2007 edition edition, sep 2007.

[145] A. Yu. Savin and B. Yu. Sternin. Noncommutative elliptic theory. Examples. Proceedings of the Steklov Institute of Mathematics, 271(1):193–211, dec 2010. 10.1134/S0081543810040152. [146] Joel L. Schiff. Normal families. Universitext. Springer-Verlag, New York, 1993. [147] Hans-Jürgen Schmeisser and Hans Triebel. Topics in and function spaces. A Wiley-Interscience Publication. John Wiley & Sons, Ltd., Chichester, 1987.

[148] Konrad Schmüdgen. Unbounded operator algebras and representation theory, volume 37 of : Advances and Applications. Birkhäuser Verlag, Basel, 1990. 10.1007/978-3-0348-7469-4. [149] Michael Schürman. Quantum q-white noise and a q-central limit theorem. Communications in Mathematical Physics, 140(3):589–615, oct 1991. 10.1007/BF02099136.

69 [150] . On the euclidean structure of relativistic field theory. Proceedings of the National Academy of Sciences of the United States of America, 44(9):956–965, sep 1958. [151] Julian Schwinger. Euclidean Quantum Electrodynamics. Physical Review, 115(3):721–731, aug 1959. Publisher: American Physical Society. 10.1103/PhysRev.115.721.

[152] I. E. Segal. A Non-Commutative Extension of Abstract Integration. Annals of Mathematics, 57(3):401–457, 1953. 10.2307/1969729.

[153] I. E. Segal. A non-commutative extension of abstract integration. Ann. of Math. (2), 57:401–457, 1953. 10.2307/1969729.

[154] I. E. Segal. Tensor Algebras Over Hilbert Spaces. II. Annals of Mathematics, 63(1):160–175, 1956. 10.2307/1969994.

[155] Irving E. Segal. Mathematical problems of relativistic physics, volume 1960 of With an appendix by George W. Mackey. Lectures in Applied Mathematics (proceedings of the Summer Seminar, Boulder, Colorado. American Mathematical Society, Providence, R.I., 1963. [156] Erhard Seiler. Schwinger functions for the Yukawa model in two dimensions with space-time cutoff. Communications in Mathematical Physics, 42(2):163–182, 1975. Publisher: Springer-Verlag.

[157] Erhard Seiler and Barry Simon. Bounds in the Yukawa2 quantum field theory: upper bound on the pressure, Hamiltonian bound and linear lower bound. Communications in Mathematical Physics, 45(2):99–114, 1975. Publisher: Springer-Verlag. [158] V. V. Shcherbakov. Elements of stochastic analysis for the case of Grassmann variables. I. Grass- mann stochastic integrals and random processes. Theoretical and Mathematical Physics, 96(1):792– 800, jul 1993. 10.1007/BF01074107. [159] V. V. Shcherbakov. Elements of stochastic analysis for the case of Grassmann variables. II. Stochas- tic partial differential equations for Grassmann processes. Theoretical and Mathematical Physics, 97(2):1229–1235, nov 1993. 10.1007/BF01016868. [160] V. V. Shcherbakov. Elements of stochastic analysis for the case of Grassmann variables. III. Correlation functions. Theoretical and Mathematical Physics, 97(3):1323–1332, dec 1993. 10.1007/BF01015761.

[161] Barry Simon. P (φ)2 Euclidean (Quantum) Field Theory. Princeton University Press, Princeton, N.J, 1974.

[162] Kalyan B. Sinha and Debashish Goswami. Quantum stochastic processes and noncommutative ge- ometry, volume 169 of Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge, 2007. 10.1017/CBO9780511618529.

[163] R. F. Streater. Classical and quantum probability. Journal of Mathematical Physics, 41(6):3556– 3603, jun 2000. 10.1063/1.533322.

[164] R. F. Streater and A. S. Wightman. PCT, spin and , and all that. Princeton Landmarks in Physics. Princeton University Press, Princeton, NJ, 2000.

[165] Franco Strocchi. An Introduction to Non-Perturbative Foundations of Quantum Field Theory. OUP Oxford, Oxford, feb 2013.

[166] Stephen J Summers. A perspective on constructive quantum field theory. ArXiv preprint arXiv:1203.3991, 2012. [167] K. Symanzik. Euclidean quantum field theory. In R. Jost, editor, Local quantum field theory (Varenna lectures), pages 152–226. Academic Press, New York, 1969. [168] François Trèves. Topological vector spaces, distributions and kernels. Academic Press, New York- London, 1967.

70 [169] Hans Triebel. Theory of function spaces. III, volume 100 of Monographs in Mathematics. Birkhäuser Verlag, Basel, 2006.

[170] Joris van der Hoeven. Majorants for formal power series. Technical Report 2003-15. Université de Paris-Sud. Département de Mathématique, Orsay, 2003.

[171] B.L. van der Waerden. Algebra: Volume I. Algebra : based in part on lectures by E. Artin and E. Noether. Springer New York, 2003.

[172] B.L. van der Waerden. Algebra: Volume II. Algebra : based in part on lectures by E. Artin and E. Noether. Springer New York, 2003.

[173] Peter van Nieuwenhuizen and Andrew Waldron. On Euclidean spinors and Wick rotations. Physics Letters B, 389(1):29–36, dec 1996. 10.1016/S0370-2693(96)01251-8. [174] J. von Neumann. On rings of operators. III. Ann. of Math. (2), 41:94–161, 1940. 10.2307/1968823. [175] Wilhelm von Waldenfels. Non-commutative algebraic central limit theorems. In Herbert Heyer, editor, Probability Measures on Groups VIII, Lecture Notes in Mathematics, pages 174–202. Berlin, Heidelberg, 1986. Springer. 10.1007/BFb0077184. [176] David N. Williams. Euclidean Fermi fields with a hermitean Feynman-Kac-Nelson formula. I. Communications in Mathematical Physics, 38(1):65–80, mar 1974. 10.1007/BF01651549. [177] Xiao Xiong, Quanhua Xu, and Zhi Yin. Sobolev, Besov and Triebel-Lizorkin Spaces on Quantum Tori, volume 252 of Memoirs of the American Mathematical Society. American Mathematical Society, mar 2018. 10.1090/memo/1203.

[178] She-Sheng Xue and Ting-chang Hsien. Stochastic quantization of fermions on lattice. Chinese Physics Letters, 2(10):474–476, oct 1985. 10.1088/0256-307X/2/10/012. [179] Wiesław Żelazko. Banach algebras. Elsevier Publishing Co., Amsterdam-London-New York; PWN– Polish Scientific Publishers, Warsaw, 1973. Translated from the Polish by Marcin E. Kuczma.

[180] J. Zinn-Justin. Quantum field theory and critical phenomena, volume 85 of International Series of Monographs on Physics. The Clarendon Press, Oxford University Press, New York, Second edition, 1993. Oxford Science Publications.

[181] Yury M. Zinoviev. Equivalence of Euclidean and Wightman field theories. Communications in Mathematical Physics, 174(1):1–27, 1995. Publisher: Springer-Verlag.

71