Abstract

The study of the multidimensional stochastic processes involves complex computations in intricate functional spaces. In particular, the diffusion processes, which include the practically important Gauss-Markov processes, are ordinarily defined through the theory of stochastic integration. Here, inspired by the Lévy-Ciesielski construction of the Wiener process, we propose an alternative representation of multidimensional Gauss-Markov processes as expansions on well-chosen Schauder bases, with independent random coefficients of normal law with zero mean and unit variance. We thereby offer a natural multiresolution description of the Gauss-Markov processes as limits of finite-dimensional partial sums of the expansion, that are strongly almost-surely convergent. Moreover, such finite-dimensional random processes constitute an optimal approximation of the process, in the sense of minimizing the associated Dirichlet energy under interpolating constraints. This approach allows for a simpler treatment of problems in many applied and theoretical fields, and we provide a short overview of applications we are currently developing.

1. Introduction

Intuitively, multidimensional continuous stochastic processes are easily conceived as solutions to randomly perturbed differential equations of the forṁ𝐗𝑡=𝑓𝐗𝑡,𝑡,𝝃𝑡,(1.1) where the perturbative term 𝝃𝑡 implicitly defines a probability space and 𝑓 satisfies some ad hoc regularity conditions. If the existence of such processes is well established for a wide range of equations through the standard Itô integration theory (see, e.g., [1]), studying their properties proves surprisingly challenging, even for the simplest multidimensional processes. Indeed, the high dimensionality of the ambient space and the nowhere differentiability of the sample paths conspire to heighten the intricacy of the sample paths spaces. In this regard, such spaces have been chiefly studied for multidimensional diffusion processes [2], and more recently, the development of rough paths theory has attracted renewed interest in the field (see [37]). However, aside from these remarkable theoretical works, little emphasis is put on the sample paths since most of the available results only make sense in distribution. This is particularly true in the Itô integration theory, where the sample path is completely neglected for the Itô map being defined up to null sets of paths.

To overcome the difficulty of working in complex multidimensional spaces, it would be advantageous to have a discrete construction of a continuous stochastic process as finite-dimensional distributions. Since we put emphasis on the description of the sample paths space, at stake is to write a process 𝐗 as an almost surely pathwise convergent series of random functions𝐗𝑡=lim𝑁𝐗𝑁𝑡with𝐗𝑁𝑡=𝑁𝑛=0𝐟𝑛(𝑡)𝚵𝑛,(1.2) where 𝐟𝑛 is a deterministic function and Ξ𝑛 is a given random variable.

The Lévy-Ciesielski construction of the 𝑑-dimensional Brownian motion 𝐖 (also referred to as Wiener process) provides us with an example of discrete representation for a continuous stochastic process. Noticing the simple form of the probability density of a Brownian bridge, it is based on completing sample paths by interpolation according to the conditional probabilities of the Wiener process [8]. More specifically, the coefficients Ξ𝑛 are Gaussian independent and the elements 𝐟𝑛, called the Schauder elements and denoted by 𝐬𝑛, are obtained by time-dependent integration of the Haar basis elements: 𝐬0,0(𝑡)=𝑡𝐈𝑑 and 𝐬𝑛,𝑘(𝑡)=𝑠𝑛,𝑘(𝑡)𝐈𝑑, with for all 𝑛>0𝑠𝑛,𝑘(𝑡)=2(𝑛1)/2𝑡𝑙𝑛,𝑘,𝑘2𝑛+1𝑡(2𝑘+1)2𝑛,2(𝑛1)/2𝑟𝑛,𝑘𝑡,(2𝑘+1)2𝑛𝑡(𝑘+1)2𝑛+1,0,otherwise.(1.3) This latter point is of relevance since, for being a Hilbert system, the introduction of the Haar basis greatly simplifies the demonstration of the existence of the Wiener process [9]. From another perspective, fundamental among discrete representations is the Karhunen-Loève decomposition giving a representation of stochastic processes by expanding it on a basis of orthogonal functions [10, 11]. The definition of the basis elements 𝑓𝑛 depends only on the second-order statistics of the considered process and the coefficients 𝜉𝑛 are pairwise uncorrelated random variables. Incidentally, such a decomposition is especially suited to study the Gaussian processes because the coefficients of the representation are Gaussian and independent. For these reasons, the Karhunen-Loève decomposition is of primary importance in exploratory data analysis, leading to methods referred to as “principal component analysis,” “Hotelling transform" [12] or “proper orthogonal decomposition” [13] according to the field of application. In particular, it was directly applied to the study of the stationary Gaussian Markov processes in the theory of random noise in radio receivers [14].

It is also important for our purpose to realize that the Schauder elements 𝐬𝑛 have compact supports that exhibit a nested structure: this fact entails that the finite sums 𝐖𝑁 are processes that interpolate the limit process 𝐖 on the endpoints of the supports, that is, on the dyadic points 𝑘2𝑁, 0𝑘2𝑁. One of the specific goal of our construction is to maintain such a property in the construction of all multidimensional the Gauss-Markov processes 𝐗 (i.e., processes that are both Gaussian and satisfy the Markov property) of the form:𝐗𝑡=𝐠(𝑡)𝑡0𝐟(𝑠)𝑑𝐖𝑠(1.4) (covering all 1-dimensional Gauss-Markov processes thanks to Doob’s representation of Gauss-Markov processes), being successively approximated by finite-dimensional processes 𝑋𝑁 that interpolates 𝐗 at ever finer resolution. In that respect, it is only in that sense that we refer to our framework as a multiresolution approach as opposed to the wavelet multiresolution theory [15]. Other multiresolution approaches have been developed for certain Gaussian processes, most notably for the fractional Brownian motion [16].

In view of this, we propose a construction of the multidimensional Gaussian Markov processes using a multiresolution Schauder basis of functions. As for the Lévy-Ciesielski construction, and in contrast with Karhunen-Loève decomposition, our basis is not made of orthogonal functions but the elements are of nested compact support and the random coefficients Ξ𝑛 are always independent and Gaussian (for convenience with law 𝒩(𝟎,𝐈𝑑), i.e., with zero mean and unitary variance). We first develop a heuristic approach for the construction of stochastic processes reminiscent of the midpoint displacement technique [8, 9], before rigorously deriving the multiresolution basis that we will be using the paper. This set of functions is then studied as a multiresolution Schauder basis of functions: in particular, we derive explicitly from the multiresolution basis an Haar-like Hilbert basis, which is the underlying structure explaining the dual relationship between basis elements and coefficients. Based on these results, we study the construction application and its inverse, the coefficient applications, that relate coefficients on the Schauder basis to sample paths. We follow up by proving the almost-sure and strong convergence of the process having independent standard normal coefficients on the Schauder basis to a Gauss-Markov process. We also show that our decomposition is optimal in some sense that is strongly evocative of spline interpolation theory [17]: the construction yields successive interpolations of the process at the interval endpoints that minimize the Dirichlet energy induced by the differential operator associated with the Gauss-Markov process [18, 19]. We also provide a series of examples for which the proposed Schauder framework yields bases of functions that have simple closed form formulae: in addition to the simple one-dimensional Markov processes, we explicit our framework for two classes of multidimensional processes, the Gauss-Markov rotations and the iteratively integrated Wiener processes (see, e.g., [2022]).

The ideas underlying this work can be directly traced back to the original work of Lévy. Here, we intend to develop a self-contained Schauder dual framework to further the description of multidimensional Gauss-Markov processes, and, in doing so, we extend some well-known results of interpolation theory in signal processing [2325]. To our knowledge, such an approach is yet to be proposed. By restraining our attention to the Gauss-Markov processes, we obviously do not assume generality. However, we hope our construction proves of interest for a number of points, which we tentatively list in the following. First, the almost-sure pathwise convergence of our construction together with the interpolation property of the finite sums allows to reformulate results of the stochastic integration in term of the geometry of finite-dimensional sample paths. In this regard, we found it appropriate to illustrate how in our framework, the Girsanov theorem for the Gauss-Markov processes appears as a direct consequence of the finite-dimensional change of variable formula. Second, the characterization of our Schauder elements as the minimizer of a Dirichlet form paves the way to the construction of infinite-dimensional Gauss-Markov processes, that is, processes whose sample points themselves are infinite-dimensional [26, 27]. Third, our construction shows that approximating a Gaussian process by a sequence of interpolating processes relies entirely on the existence of a regular triangularization of the covariance operator, suggesting to further investigate this property for non-Markov Gaussian processes [28]. Finally, there is a number of practical applications where applying the Schauder basis framework clearly provides an advantage compared to standard stochastic calculus methods, among which first-hitting times of stochastic processes, pricing of multidimensional path-dependant options [2932], regularization technique for support vector machine learning [33], and more theoretical work on uncovering the differential geometry structure of the space of the Gauss-Markov stochastic processes [34]. We conclude our exposition by developing in more detail some of these direct implications which will be the subjects of forthcoming papers.

2. Heuristic Approach to the Construction

In order to provide a discrete multiresolution description of the Gauss-Markov processes, we first establish basic results about the law of the Gauss-Markov bridges in the multidimensional setting. We then use them to infer the candidate expressions for our desired bases of functions, while imposing its elements to be compactly supported on nested sequence segments. Throughout this paper, we are working in a complete probability space (Ω,,𝐏).

2.1. Multidimensional Gauss-Markov Processes

After recalling the definition of the multidimensional Gauss-Markov processes in terms of stochastic integral, we use the well-known conditioning formula for the Gaussian vectors to characterize the law of the Gauss-Markov bridge processes.

2.1.1. Notations and Definitions

Let (𝐖𝑡,𝑡,𝑡[0,1]) be an 𝑚-dimensional Wiener process, consider the continuous functions 𝜶[0,1]𝑑×𝑑Γ[0,1]𝑑×𝑚, and define the positive bounded continuous function Γ=ΓΓ𝑇[0,1]𝑑×𝑑. The 𝑑-dimensional Ornstein-Uhlenbeck process associated with these parameters is solution of the equation𝑑𝐗𝑡=𝜶(𝑡)𝐗𝑡𝑑𝑡+𝚪(𝑡)𝑑𝐖𝑡,(2.1) and with initial condition 𝐗𝑡0 in 𝑡0, it reads𝐗𝑡=𝐅𝑡0,𝑡𝐗𝑡0+𝐅𝑡0,𝑡𝑡𝑡0𝐅𝑠,𝑡0𝚪(𝑠)𝑑𝐖𝑠,(2.2) where 𝐅(𝑡0,𝑡) is the flow of the equation, namely, the solution in 𝑑×𝑑 of the linear equation:𝜕𝐅𝑡0,𝑡𝜕𝑡=𝜶(𝑡)𝐅(𝑡),𝐅𝑡0,𝑡0=𝐈𝑑.(2.3) Note that the flow 𝐅(𝑡0,𝑡) enjoys the chain rule property:𝐅𝑡0,𝑡=𝐅𝑡1,𝑡𝐅𝑡0,𝑡1.(2.4) For all 𝑡,𝑠 such that 𝑡0<𝑠,𝑡, the vectors 𝐗𝑡 and 𝐗𝑠 admit the covariance𝐂𝑡0(𝑠,𝑡)=𝐅𝑡0,𝑡𝑡𝑠𝑡0𝐅𝑤,𝑡0𝚪(𝑤)𝐅𝑤,𝑡0𝑇𝑑𝑤𝐅𝑡0,𝑠𝑇=𝐅𝑡0,𝑡𝐡𝑡0(𝑠,𝑡)𝐅𝑡0,𝑠𝑇,(2.5) where we further defined 𝐡𝑢(𝑠,𝑡) the function𝐡𝑢(𝑠,𝑡)=𝑡𝑠𝐅(𝑤,𝑢)𝚪(𝑤)𝐅(𝑤,𝑢)𝑇𝑑𝑤,(2.6) which will be of particular interest in the sequel. Note that because of the chain rule property of the flow, we have 𝐡𝑣(𝑠,𝑡)=𝐅(𝑣,𝑢)𝐡𝑢(𝑠,𝑡)𝐅(𝑣,𝑢)𝑇.(2.7) We suppose that the process 𝐗 is never degenerated, that is, for all 𝑡0<𝑢<𝑣, all the components of the vector 𝐗𝑣 taking into account 𝐗𝑢 are nondeterministic random variables, which is equivalent to saying that the covariance matrix of 𝐗𝑣 taking into account 𝐗𝑢, denoted by 𝐂𝑢(𝑣,𝑣) is symmetric positive definite for any 𝑢𝑣. Therefore, assuming the initial condition 𝐗0=𝟎, the multidimensional centered process 𝐗 has a representation (similar to Doob’s representation for one-dimensional processes, see [35]) of form𝐗𝑡=𝐠(𝑡)𝑡0𝐟(𝑠)𝑑𝐖𝑠,(2.8) with 𝐠(𝑡)=𝐅(0,𝑡) and 𝐟(𝑡)=𝐅(𝑡,0)Γ(𝑡).

Note that the processes considered in this paper are defined on the time interval [0,1]. However, because of the time-rescaling property of these processes, considering the processes on this time interval is equivalent to considering the process on any other bounded interval without loss of generality.

2.1.2. Conditional Law and Gauss-Markov Bridges

As stated in the introduction, we aim at defining a multiresolution description of Gauss-Markov processes. Such a description can be seen as a multiresolution interpolation of the process that is getting increasingly finer. This principle, in addition to the Markov property, prescribes to characterize the law of the corresponding Gauss-Markov bridge, that is, the Gauss-Markov process under consideration, conditioned on its initial and final values. The bridge process of the Gauss process is still a Gauss process and, for a Markov process, its law can be computed as follows.

Proposition 2.1. Let 𝑡𝑥𝑡𝑧 two times in the interval [0,1]. For any 𝑡[𝑡𝑥,𝑡𝑧], the random variable 𝐗𝑡 conditioned on 𝐗𝑡𝑥=𝐱 and 𝐗𝑡𝑧=𝐳 is a Gaussian variable with covariance matrix 𝚺(𝑡) and mean vector 𝝁(𝑡) given by 𝚺𝑡;𝑡𝑥,𝑡𝑧=𝐡𝑡𝑡𝑥,𝑡𝐡𝑡(𝑡𝑥,𝑡𝑧)1𝐡𝑡𝑡,𝑡𝑧,𝝁(𝑡)=𝝁𝑙𝑡;𝑡𝑥,𝑡𝑧𝐱+𝝁𝑟𝑡;𝑡𝑥,𝑡𝑧𝐳,(2.9) where the continuous matrix functions 𝝁𝑙(;𝑡𝑥,𝑡𝑧) and 𝝁𝑟(;𝑡𝑥,𝑡𝑧) of 𝑑×𝑑 are given by 𝝁𝑙𝑡;𝑡𝑥,𝑡𝑧=𝐅𝑡𝑥,𝑡𝐡𝑡𝑥𝑡,𝑡𝑧𝐡𝑡𝑥𝑡𝑥,𝑡𝑧1,𝝁𝑟𝑡;𝑡𝑥,𝑡𝑧=𝐅𝑡𝑧,𝑡𝐡𝑡𝑧𝑡𝑥,𝑡𝐡𝑡𝑧(𝑡𝑥,𝑡𝑧)1.(2.10)

Note that the functions 𝜇𝑙 and 𝜇𝑟 have the property that 𝝁𝑙(𝑡𝑥;𝑡𝑥,𝑡𝑧)=𝝁𝑟(𝑡𝑧;𝑡𝑥,𝑡𝑧)=𝐈𝑑 and 𝝁𝑙(𝑡𝑧;𝑡𝑥,𝑡𝑧)=𝝁𝑟(𝑡𝑥;𝑡𝑥,𝑡𝑧)=𝟎 ensuring that the process is indeed equal to 𝐱 at time 𝑡𝑥 and 𝐳 at time 𝑡𝑧.

Proof. Let 𝑡𝑥,𝑡𝑧 be two times of the interval [0,1] such that 𝑡𝑥<𝑡𝑧, and let 𝑡[𝑡𝑥,𝑡𝑧]. We consider the Gaussian random variable 𝝃=(𝐗𝑡,𝐗𝑡𝑧) conditioned on the fact that 𝐗𝑡𝑥=𝐱. Its mean can be easily computed from expression (2.2) and reads 𝐦𝑡,𝐦𝑡𝑧=𝐅𝑡𝑥,𝑡𝐱,𝐅𝑡𝑥,𝑡𝑧𝐱=𝐠(𝑡)𝐠1𝑡𝑥𝐱,𝐠𝑡𝑧𝐠1𝑡𝑥𝐱,(2.11) and its covariance matrix, from (2.5), reads 𝐂𝑡,𝑡𝐂𝑡,𝑡𝑧𝐂𝑡𝑧,𝑡𝐂𝑡𝑧,𝑡𝑧=𝐅𝑡𝑥,𝑡𝐡𝑡𝑥𝑡𝑥,𝑡𝐅𝑡𝑥,𝑡𝑇𝐅𝑡𝑥,𝑡𝐡𝑡𝑥𝑡𝑥,𝑡𝐅𝑡𝑥,𝑡𝑧𝑇𝐅𝑡𝑥,𝑡𝑧𝐡𝑡𝑥𝑡𝑥,𝑡𝐅𝑡𝑥,𝑡𝑇𝐅𝑡𝑥,𝑡𝑧𝐡𝑡𝑥𝑡𝑥,𝑡𝑧𝐅𝑡𝑥,𝑡𝑧𝑇=𝐡𝑡𝑡𝑥,𝑡𝐡𝑡𝑡𝑥,𝑡𝐅𝑡,𝑡𝑧𝑇𝐅𝑡,𝑡𝑧𝐡𝑡𝑡𝑥,𝑡𝐅𝑡,𝑡𝑧𝐡𝑡𝑡𝑥,𝑡𝑧𝐅𝑡,𝑡𝑧𝑇.(2.12) From there, we apply the conditioning formula for the Gaussian vectors (see, e.g., [36]) to infer the law of 𝐗𝑡 conditioned on 𝐗𝑡𝑥=𝐱 and 𝐗𝑡𝑧=𝐳, that is the law 𝒩(𝝁(𝑡),𝚺(𝑡;𝑡𝑥,𝑡𝑧)) of 𝐁𝑡 where 𝐁 denotes the bridge process obtained by pinning 𝐗 in 𝑡𝑥 and 𝑡𝑧. The covariance matrix is given by 𝚺𝑡;𝑡𝑥,𝑡𝑧=𝐂𝑦,𝑦𝐂𝑦,𝑧𝐂1𝑧,𝑧𝐂𝑧,𝑦=𝐡𝑡𝑡𝑥,𝑡𝐡𝑡𝑡𝑥,𝑡𝐡𝑡𝑡𝑥,𝑡𝑧1𝐡𝑡𝑡𝑥,𝑡=𝐡𝑡𝑡𝑥,𝑡𝐡𝑡𝑡𝑥,𝑡𝑧1𝐡𝑡𝑡,𝑡𝑧,(2.13) and the mean reads 𝝁(𝑡)=𝐦𝑦+𝐂𝑦,𝑧𝐂1𝑧,𝑧𝐳𝐦𝑧=𝐅𝑡𝑥,𝑡𝐈𝑑𝐡𝑡𝑥𝑡𝑥,𝑡𝐡𝑡𝑥(𝑡𝑥,𝑡𝑧)1𝐱+𝐅𝑡𝑧,𝑡𝐡𝑡𝑧𝑡𝑥,𝑡𝐡𝑡𝑧(𝑡𝑥,𝑡𝑧)1𝐳,=𝐅(𝑡𝑥,𝑡)𝐡𝑡𝑥(𝑡,𝑡𝑧)𝐡𝑡𝑥(𝑡𝑥,𝑡𝑧)1𝝁𝑙(𝑡;𝑡𝑥,𝑡𝑧)𝐱+𝐅(𝑡𝑧,𝑡)𝐡𝑡𝑧(𝑡𝑥,𝑡)𝐡𝑡𝑧(𝑡𝑥,𝑡𝑧)1𝝁𝑟(𝑡;𝑡𝑥,𝑡𝑧)𝐳,(2.14) where we have used the fact that 𝐡𝑡𝑥(𝑡𝑥,𝑡𝑧)=𝐡𝑡𝑧(𝑡𝑥,𝑡)+𝐡𝑡𝑥(𝑡,𝑡𝑧). The regularity of the thus-defined functions 𝝁𝑥 and 𝝁𝑧 directly stems from the regularity of the flow operator 𝐅. Moreover, since for any 0𝑡,𝑢1, we observe that 𝐅(𝑡,𝑡)=𝐈𝑑 and 𝑢(𝑡,𝑡)=𝟎; we clearly have 𝝁𝑥(𝑡𝑥)=𝝁𝑦(𝑡)=𝐈𝑑 and 𝝁𝑥(𝑡)=𝝁𝑦(𝑡𝑥)=0.

Remark 2.2. Note that these laws can also be computed using the expression of the density of the processes but involve more intricate calculations. An alternative approach also provides a representation of Gauss-Markov bridges with the use of integral and anticipative representation [37]. These approaches allow to compute the probability distribution of the Gauss-Markov bridge as a process (i.e., allows to compute the covariances), but since this will be of no use in the sequel, we do not provide the expressions.

2.2. The Multiresolution Description of Gauss-Markov Processes

Recognizing the Gauss property and the Markov property as the two crucial elements for a stochastic process to be expanded to Lévy-Cesielski, our approach first proposes to exhibit bases of deterministic functions that would play the role of the Schauder bases for the Wiener process. In this regard, we first expect such functions to be continuous and compactly supported on increasingly finer supports (i.e., subintervals of the definition interval [0,1]) in a similar nested binary tree structure. Then, as in the Lévy-Ciesielski construction, we envision that, at each resolution (i.e., on each support), the partially constructed process (up to the resolution of the support) has the same conditional expectation as the Gauss-Markov process when conditioned on the endpoints of the supports. The partial sums obtained with independent Gaussian coefficients of law 𝒩(0,1) will thus approximate the targeted Gauss-Markov process in a multiresolution fashion, in the sense that, at every resolution, considering these two processes on the interval endpoints yields finite-dimensional Gaussian vectors of the same law.

2.2.1. Nested Structure of the Sequence of Supports

Here, we define the nested sequence of segments that constitute the supports of the multiresolution basis. We construct such a sequence by recursively partitioning the interval [0,1].

More precisely, starting from 𝑆1,0=[𝑙1,0,𝑟1,0] with 𝑙1,0=0 and 𝑟1,0=1, we iteratively apply the following operation. Suppose that, at the 𝑛th step, the interval [0,1] is decomposed into 2𝑛1 intervals 𝑆𝑛,𝑘=[𝑙𝑛,𝑘,𝑟𝑛,𝑘], called supports, such that 𝑙𝑛,𝑘+1=𝑟𝑛,𝑘 for 0𝑘<2𝑛1. Each of these intervals is then subdivided into two child intervals, a left-child 𝑆𝑛+1,2𝑘 and a right-child 𝑆𝑛+1,2𝑘+1, and the subdivision point 𝑟𝑛+1,2𝑘=𝑙𝑛+1,2𝑘+1 is denoted by 𝑚𝑛,𝑘. Therefore, we have defined three sequences of real 𝑙𝑛,𝑘, 𝑚𝑛,𝑘, and 𝑟𝑛,𝑘 for 𝑛>0 and 0𝑘<2𝑛1 satisfying 𝑙0,0=0𝑙𝑛,𝑘<𝑚𝑛,𝑘<𝑟𝑛,𝑘𝑟0,0=1 and𝑙𝑛+1,2𝑘=𝑙𝑛,𝑘,𝑚𝑛,𝑘=𝑟𝑛+1,2𝑘=𝑙𝑛+1,2𝑘+1,𝑟𝑛+1,2𝑘+1=𝑟𝑛,𝑘(2.15) with the convention 𝑙0,0=0 and 𝑟0,0=1 and 𝑆0,0=[0,1]. The resulting sequence of supports {𝑆𝑛,𝑘;𝑛0,0𝑘<2𝑛1} clearly has a binary tree structure.

For the sake of compactness of notations, we define the set of indices=𝑛<𝑁𝑛with𝑁=(𝑛,𝑘)20<𝑛𝑁,0𝑘<2𝑛1,(2.16) and for 𝑁>0, we define 𝐷𝑁={𝑚𝑛,𝑘,(𝑛,𝑘)𝑁1}{0,1}, the set of endpoints of the intervals 𝑆𝑁,𝑘. We additionally require that there exists 𝜌(0,1) such that for all (𝑛,𝑘)max(𝑟𝑛,𝑘𝑚𝑛,𝑘,𝑚𝑛,𝑘𝑙𝑛,𝑘)<𝜌(𝑟𝑛,𝑘𝑙𝑛,𝑘) which in particular implies thatlim𝑛sup𝑘𝑟𝑛,𝑘𝑙𝑛,𝑘=0(2.17) and ensures that the set of endpoints 𝑁N𝐷𝑁 is everywhere dense in [0,1]. The simplest case of such partitions is the dyadic partition of [0,1], where the endpoints for (𝑛,𝑘) read𝑙𝑛,𝑘=𝑘2𝑛+1,𝑚𝑛,𝑘=(2𝑘+1)2𝑛,𝑟𝑛,𝑘=(𝑘+1)2𝑛+1,(2.18) in which case the endpoints are simply the dyadic points 𝑁𝐷𝑁={𝑘2𝑁|0𝑘2𝑁}. Figure 1 represents the global architecture of the nested sequence of intervals.

The nested structure of the supports, together with the constraint of continuity of the bases elements, implies that only a finite number of coefficients are needed to construct the exact value of the process at a given endpoint, thus providing us with an exact schema to simulate the sample values of the process on the endpoint up to an arbitrary resolution, as we will further explore.

2.2.2. Innovation Processes for Gauss-Markov Processes

For 𝐗𝑡, a multidimensional Gauss-Markov process, we call the multiresolution description of a process the sequence of conditional expectations on the nested sets of endpoints 𝐷𝑛. In detail, if we denote by 𝑁 the filtration generated by {𝐗𝑡;𝑡𝐷𝑁} given the values of the process at the endpoints 𝐷𝑁 of the partition, we introduce the sequence of the Gaussian processes (𝐙𝑁𝑡)𝑁1 defined by:𝐙𝑁𝑡=𝔼𝐗𝑡𝑁=𝔼𝑁𝐗𝑡.(2.19) These processes 𝐙𝑁 can be naturally viewed as an interpolation of the process 𝐗 sampled at the increasingly finer times 𝐷𝑁 since for all 𝑡𝐷𝑁 we have 𝐙𝑁𝑡=𝐗𝑁𝑡. The innovation process (𝜹𝑁𝑡,𝑡,𝑡[0,1]) is defined as the update transforming the process 𝐙𝑁𝑡 into 𝐙𝑁+1𝑡, that is, 𝜹𝑁𝑡=𝐙𝑁+1𝑡𝐙𝑁𝑡.(2.20) It corresponds to the difference the additional knowledge of the process at the points 𝑚𝑁,𝑘 make on the conditional expectation of the process. This process satisfies the following important properties that found our multiresolution construction.

Proposition 2.3. The innovation process 𝜹𝑁𝑡 is a centered Gaussian process independent of the processes 𝐙𝑛𝑡 for any 𝑛𝑁. For 𝑠𝑆𝑁,𝑘 and 𝑡𝑆𝑁,𝑝 with 𝑘,𝑝𝑁, the covariance of the innovation process reads 𝔼𝑁𝜹𝑁𝑡𝜹𝑁𝑠𝑇=𝝁𝑁,𝑘(𝑡)𝚺𝑁,𝑘𝝁𝑁,𝑘(𝑡)𝑇if𝑘=𝑝,𝟎if𝑘𝑝,(2.21) where 𝝁𝑁,𝑘(𝑡)=𝝁𝑟𝑡;𝑙𝑁,𝑘,𝑚𝑁,𝑘,𝑡𝑙𝑁,𝑘,𝑚𝑁,𝑘,𝝁𝑙𝑡;𝑚𝑁,𝑘,𝑟𝑁,𝑘,𝑡𝑚𝑁,𝑘,𝑟𝑁,𝑘.(2.22) with 𝝁𝑙, 𝝁𝑟 and 𝚺𝑁,𝑘=𝚺(𝑚𝑁,𝑘;𝑙𝑁,𝑘,𝑟𝑁,𝑘) as defined in Proposition 2.1.

Proof. Because of the Markovian property of the process 𝐗, the law of the process 𝐙𝑁 can be computed from the bridge formula derived in Proposition 2.1 and we have 𝐙𝑁𝑡=𝝁𝑙𝑡;𝑙𝑁,𝑘,𝑟𝑁,𝑘𝐗𝑙𝑁,𝑘+𝝁𝑟𝑡;𝑙𝑁,𝑘,𝑟𝑁,𝑘𝐗𝑟𝑁,𝑘,𝐙𝑁+1𝑡=𝝁𝑙𝑡;𝑙𝑁,𝑘,𝑚𝑁,𝑘𝐗𝑙𝑁,𝑘+𝝁𝑟𝑡;𝑙𝑁,𝑘,𝑚𝑁,𝑘𝐗𝑚𝑁,𝑘,for𝑡𝑙𝑁,𝑘,𝑚𝑁,𝑘,𝝁𝑙𝑡;𝑚𝑁,𝑘,𝑟𝑁,𝑘𝐗𝑚𝑁,𝑘+𝝁𝑟𝑡;𝑚𝑁,𝑘,𝑟𝑁,𝑘𝐗𝑟𝑁,𝑘,for𝑡𝑙𝑁,𝑘,𝑚𝑁,𝑘.(2.23) Therefore, the innovation process can be written for 𝑡𝑆𝑁,𝑘 as 𝜹𝑁𝑡=𝝁𝑁𝑁,𝑘(𝑡)𝐗𝑚𝑁,𝑘+𝝂𝑁(𝑡)𝐐𝑁𝑡,(2.24) where 𝐐𝑁𝑡 is a 𝑁 measurable process 𝝂𝑁(𝑡) a deterministic matrix function and 𝝁𝑁,𝑘(𝑡)=𝝁𝑟𝑡;𝑙𝑁,𝑘,𝑚𝑁,𝑘,𝑡𝑙𝑁,𝑘,𝑚𝑁,𝑘,𝝁𝑙𝑡;𝑚𝑁,𝑘,𝑟𝑁,𝑘,𝑡𝑚𝑁,𝑘,𝑟𝑁,𝑘.(2.25) The expressions of 𝝂 and 𝐐 are quite complex but are highly simplified when noting that 𝔼𝜹𝑁𝑡𝑁=𝔼𝐙𝑁+1𝑡𝑁𝐙𝑁𝑡=𝔼𝔼𝐙𝑡𝑁+1𝑁𝐙𝑁𝑡=𝟎(2.26) directly implying that 𝝂(𝑡)𝐐𝑁𝑡=𝝁𝑁(𝑡)𝐙𝑁𝑚𝑁,𝑘 and yielding the remarkably compact expression 𝜹𝑁𝑡=𝝁𝑁,𝑘(𝑡)𝐗𝑚𝑁,𝑘𝐙𝑁𝑚𝑁,𝑘.(2.27) This process is a centered Gaussian process. Moreover, observing that it is 𝑁-measurable, it can be written as 𝜹𝑁𝑡=𝝁𝑁,𝑘(𝑡)𝐗𝑚𝑁,𝑘𝑁𝐙𝑁𝑚𝑁,𝑘,(2.28) and the process {𝐗𝑚𝑁,𝑘𝑁} appears as the Gauss-Markov bridge conditioned at times 𝑙𝑁,𝑘 and 𝑟𝑁,𝑘, and whose covariance is given by Proposition 2.1 and that has the expression 𝚺𝑁,𝑘=𝚺𝑚𝑁,𝑘;𝑙𝑁,𝑘,𝑟𝑁,𝑘=𝐡𝑚𝑛,𝑘𝑙𝑛,𝑘,𝑚𝑛,𝑘𝐡𝑚𝑛,𝑘𝑙𝑛,𝑘,𝑟𝑛,𝑘1𝐡𝑚𝑛,𝑘𝑚𝑛,𝑘,𝑟𝑛,𝑘.(2.29) Let (𝑠,𝑡)[0,1]2, and assume that 𝑠𝑆𝑁,𝑘 and 𝑡𝑆𝑁,𝑝. If 𝑘𝑝, then, because of the Markov property of the process 𝐗, the two bridges are independent and therefore the covariance 𝔼𝑁[𝜹𝑁𝑡(𝜹𝑁𝑠)𝑇] is zero. If 𝑘=𝑝, we have 𝔼𝑁𝜹𝑁𝑡𝜹𝑁𝑠𝑇=𝝁𝑁,𝑘(𝑡)𝚺𝑁,𝑘𝝁𝑁,𝑘(𝑠)𝑇.(2.30) Eventually, the independence property stems from the simple properties of the conditional expectation. Indeed, let 𝑛𝑁. We have 𝔼𝐙𝑛𝑡𝜹𝑁𝑠𝑇=𝔼𝐙𝑛𝑡𝐙𝑁+1𝑠𝐙𝑁𝑠𝑇=𝔼𝔼𝐗𝑡𝑛𝔼𝐗𝑇𝑠𝑁+1𝔼𝐗𝑇𝑠𝑁=𝔼𝔼𝐗𝑡𝑛𝔼𝐗𝑇𝑠𝑁+1𝔼𝔼𝐗𝑡𝑛𝔼𝐗𝑇𝑠𝑁=𝔼𝐙𝑛𝑡𝐙𝑛𝑠𝑇𝔼𝐙𝑛𝑡𝐙𝑛𝑠𝑇=𝟎(2.31) and the fact that a zero covariance between two Gaussian processes implies the independence of these processes concludes the proof.

2.2.3. Derivation of the Candidate Multiresolution Bases of Functions

We deduce from the previous proposition the following fundamental theorem of this paper.

Theorem 2.4. For all 𝑁, there exists a collection of 𝝍𝑁,𝑘[0,1]𝑑×𝑑 that are zero outside the subinterval 𝑆𝑁,𝑘 such that in distribution one has: 𝜹𝑁𝑡=𝑘𝑁𝝍𝑁,𝑘(𝑡)𝚵𝑁,𝑘,(2.32) where Ξ𝑁,𝑘 are independent 𝑑-dimensional standard normal random variables (i.e., of law 𝒩(0,𝐈𝑑)). This basis of functions is unique up to an orthogonal transformation.

Proof. The two processes 𝜹𝑁𝑡 and 𝐝𝑁𝑡def=𝑘𝑁𝝍𝑁,𝑘(𝑡)Ξ𝑁,𝑘 are two Gaussian processes of mean zero. Therefore, we are searching for functions 𝝍𝑁,𝑘 vanishing outside 𝑆𝑁,𝑘 and ensuring that the two processes have the same probability distribution. A necessary and sufficient condition for the two processes to have the same probability distribution is to have the same covariance function (see, e.g., [36]). We therefore need to show the existence of a collection of functions 𝝍𝑁,𝑘(𝑡) functions that vanish outside the subinterval 𝑆𝑁,𝑘 and that ensure that the covariance of the process 𝐝𝑁 is equal to the covariance of 𝜹𝑁. Let (𝑠,𝑡)[0,1] such that 𝑠𝑆𝑁,𝑘 and 𝑡𝑆𝑁,𝑝. If 𝑘𝑝, the assumption fact that the functions 𝜓𝑁,𝑘 vanish outside 𝑆𝑁,𝑘 implies that 𝔼𝐝𝑁𝑡𝐝𝑁𝑠𝑇=𝟎.(2.33) If 𝑘=𝑝, the covariance reads 𝔼𝐝𝑁𝑡𝐝𝑁𝑠𝑇=𝔼𝝍𝑁,𝑘(𝑡)𝚵𝑁,𝑘𝚵𝑇𝑁,𝑘𝝍𝑁,𝑘(𝑠)𝑇=𝝍𝑁,𝑘(𝑡)𝝍𝑁,𝑘(𝑠)𝑇,(2.34) which needs to be equal to the covariance of 𝜹𝑁, namely, 𝝍𝑁,𝑘(𝑡)𝝍𝑁,𝑘(𝑠)𝑇=𝝁𝑁,𝑘(𝑡)𝚺𝑁,𝑘𝝁𝑁,𝑘(𝑠)𝑇.(2.35) Therefore, since 𝝁𝑁,𝑘(𝑚𝑁,𝑘)=𝐈𝑑, we have 𝝍𝑁,𝑘𝑚𝑁,𝑘𝝍𝑁,𝑘𝑚𝑁,𝑘𝑇=𝚺𝐍,𝐤.(2.36) We can hence now define 𝝍𝑁,𝑘(𝑚𝑁,𝑘) as a square root 𝝈𝑁,𝑘 of the symmetric positive matrix 𝚺𝐍,𝐤, by fixing 𝑠=𝑚𝑁,𝑘 in (2.35) 𝝍𝑁,𝑘(𝑡)𝝈𝑇𝑁,𝑘=𝝁(𝑡)𝝈𝑁,𝑘𝝈𝑇𝑁,𝑘.(2.37) Eventually, since by assumption we have that 𝚺𝑁,𝑘 is invertible, so is 𝝈𝑁,𝑘, and the functions 𝝍𝑁,𝑘 can be written as 𝝍𝑁,𝑘(𝑡)=𝝁𝑁,𝑘(𝑡)𝝈𝑁,𝑘(2.38) with 𝝈𝑁,𝑘 being a square root of 𝚺𝑁,𝑘. Square roots of positive symmetric matrices are uniquely defined up to an orthogonal transformation. Therefore, all square roots of 𝚺𝑁,𝑘 are related by orthogonal transformations 𝝈𝑁,𝑘=𝝈𝑁,𝑘𝐎𝑁,𝑘, where 𝐎𝑁,𝑘𝐎𝑇𝑁,𝑘=𝐈𝑑. This property immediately extends to the functions 𝝍𝑁,𝑘 we are studying: two different functions 𝝍𝑁,𝑘 and 𝝍𝑁,𝑘 satisfying the theorem differ from an orthogonal transformation 𝐎𝑁,𝑘. We proved that, for 𝝍𝑁,𝑘(𝑡)Ξ𝑁,𝑘 to have the same law as 𝜹𝑁(𝑡) in the interval 𝑆𝑁,𝑘, the functions 𝝍𝑁,𝑘 with support in 𝑆𝑁,𝑘 are necessarily of the form 𝝁𝑁,𝑘(𝑡)𝝈𝑁,𝑘. It is straightforward to show the sufficient condition that provided such a set of functions, the processes 𝜹𝑁𝑡 and 𝐝𝑁𝑡 are equal in law, which ends the proof of the theorem.

Using the expressions obtained in Proposition 2.1, we can make completely explicit the form of the basis in terms of the functions 𝐟,𝐠, and 𝐡:𝝍𝑛,𝑘(𝑡)=𝐠(𝑡)𝐠1𝑚𝑛,𝑘𝐡𝑚𝑛,𝑘𝑙𝑛,𝑘,𝑡𝐡𝑚𝑛,𝑘𝑙𝑛,𝑘,𝑚𝑛,𝑘1𝝈𝑛,𝑘,for𝑙𝑛,𝑘𝑡𝑚𝑛,𝑘,𝐠(𝑡)𝐠1𝑚𝑛,𝑘𝐡𝑚𝑛,𝑘𝑡,𝑟𝑛,𝑘𝐡𝑚𝑛,𝑘𝑚𝑛,𝑘,𝑟𝑛,𝑘1𝝈𝑛,𝑘,for𝑚𝑛,𝑘𝑡𝑟𝑛,𝑘,(2.39) and 𝝈𝑛,𝑘 satisfies𝝈𝑛,𝑘𝝈𝑇𝑛,𝑘=𝐡𝑚𝑛,𝑘𝑙𝑛,𝑘,𝑚𝑛,𝑘𝐡𝑚𝑛,𝑘𝑙𝑛,𝑘,𝑟𝑛,𝑘1𝐡𝑚𝑛,𝑘𝑚𝑛,𝑘,𝑟𝑛,𝑘.(2.40) Note that 𝝈𝑛,𝑘 can be defined uniquely as the symmetric positive square root, or as the lower triangular matrix resulting from the Cholesky decomposition of 𝚺𝑛,𝑘.

Let us now define the function 𝝍0,0[0,1]𝑑×𝑑 such that the process 𝝍0,0(𝑡)Ξ0,0 has the same covariance as 𝐙0𝑡, which is computed using exactly the same technique as that developed in the proof of Theorem 2.4 and that has the expression𝝍0,0(𝑡)=𝐠(𝑡)𝐡0𝑙0,0,𝑡𝐡0(𝑙0,0,𝑟0,0)1𝐠1𝑟0,0𝝈0,0,(2.41) for 𝝈0,0, a square root of 𝐂𝑟0,0 the covariance matrix of 𝐗𝑟0,0 which from (2.5) reads𝐅(0,1)𝐡0(1,1)𝐅(0,1)𝑇=𝐠(1)𝐡0(1,1)(𝐠(1))𝑇.(2.42)

We are now in position to show the following corollary of Theorem 2.4.

Corollary 2.5. The Gauss-Markov process 𝐙𝑁𝑡 is equal in law to the process 𝐗𝑁𝑡=𝑁1𝑛=0𝑘𝑛𝝍𝑛,𝑘(𝑡)𝚵𝑛,𝑘,(2.43) where Ξ𝑛,𝑘 are independent standard normal random variables 𝒩(0,𝐈𝑑).

Proof. We have 𝐙𝑁𝑡=𝐙𝑁𝑡𝐙𝑁1𝑡+𝐙𝑁1𝑡𝐙𝑁2𝑡++𝐙2𝑡𝐙1𝑡+𝐙1𝑡=𝑁1𝑛=1𝜹𝑛𝑡+𝐙1𝑡=𝑁1𝑛=1𝑘𝑛𝝍𝑛,𝑘(𝑡)𝚵𝑛,𝑘+𝝍0,0(𝑡)𝚵0,0=𝑁1𝑛=0𝑘𝑛𝝍𝑛,𝑘(𝑡)𝚵𝑛,𝑘.(2.44)

We therefore identified a collection of functions {𝝍𝑛,𝑘}(𝑛,𝑘) that allows a simple construction of the Gauss-Markov process iteratively conditioned on increasingly finer partitions of the interval [0,1]. We will show that this sequence 𝐙𝑁𝑡 converges almost surely towards the Gauss-Markov process 𝐗𝑡 used to construct the basis, proving that these finite-dimensional continuous functions 𝐙𝑁𝑡 form an asymptotically accurate description of the initial process. Beforehand, we rigorously study the Hilbertian properties of the collection of functions we just defined.

3. The Multiresolution Schauder Basis Framework

The above analysis motivates the introduction of a set of functions {𝝍𝑛,𝑘}(𝑛,𝑘) we now study in details. In particular, we enlighten the structure of the collection of functions 𝝍𝑛,𝑘 as a Schauder basis in a certain space 𝒳 of continuous functions from [0,1] to 𝑑. The Schauder structure was defined in [38, 39], and its essential characterization is the unique decomposition property: namely that every element 𝑥 in 𝒳 can be written as a well-formed linear combination𝑥=(𝑛,𝑘)𝝍𝑛,𝑘𝝃𝑛,𝑘,(3.1) and that the coefficients satisfying the previous relation are unique.

3.1. System of Dual Bases

To complete this program, we need to introduce some quantities that will play a crucial role in expressing the family 𝝍𝑛,𝑘 as a Schauder basis for some given space. In (2.39), two constant matrices 𝑑×𝑑 appear that will have a particular importance in the sequel for (𝑛,𝑘) in with 𝑛0:𝐋𝑛,𝑘=𝐠𝑇𝑚𝑛,𝑘𝐡𝑚𝑛,𝑘𝑙𝑛,𝑘,𝑚𝑛,𝑘1𝝈𝑛,𝑘=𝐡𝑙𝑛,𝑘,𝑚𝑛,𝑘1𝐠1𝑚𝑛,𝑘𝝈𝑛,𝑘,𝐑𝑛,𝑘=𝐠𝑇𝑚𝑛,𝑘𝐡𝑚𝑛,𝑘𝑚𝑛,𝑘,𝑟𝑛,𝑘1𝝈𝑛,𝑘=𝐡𝑚𝑛,𝑘,𝑟𝑛,𝑘1𝐠1𝑚𝑛,𝑘𝝈𝑛,𝑘,(3.2) where 𝐡 stands for 𝐡0. We further define the matrix𝐌𝑛,𝑘=𝐠𝑇𝑚𝑛,𝑘𝝈1𝑛,𝑘𝑇(3.3) and we recall that 𝝈𝑛,𝑘 is a square root of 𝚺𝑛,𝑘, the covariance matrix of 𝐗𝑚𝑛,𝑘, conditionally to 𝐗𝑙𝑛,𝑘 and 𝐗𝑟𝑛,𝑘, given in (2.29). We stress that the matrices 𝐋𝑛,𝑘,𝐑𝑛,𝑘, 𝐌𝑛,𝑘, and 𝚺𝑛,k are all invertible and satisfy the important following properties.

Proposition 3.1. For all (𝑛,𝑘) in , 𝑛0, one has:(i)𝐌𝑛,𝑘=𝐋𝑛,𝑘+𝐑𝑛,𝑘(ii)𝚺1𝑛,𝑘=(𝐡𝑚𝑛,𝑘(𝑙𝑛,𝑘,𝑚𝑛,𝑘))1+(𝐡𝑚𝑛,𝑘(𝑚𝑛,𝑘,𝑟𝑛,𝑘))1.

To prove this proposition, we first establish the following simple lemma of linear algebra.

Lemma 3.2. Given two invertible matrices 𝐴 and 𝐵 in 𝐺𝐿𝑛() such that 𝐶=𝐴+𝐵 is also invertible, if one defines 𝐷=𝐴𝐶1𝐵, one has the following properties:(i)𝐷=𝐴𝐶1𝐵=𝐵𝐶1𝐴(ii)𝐷1=𝐴1+𝐵1.

Proof. (i)𝐷=𝐴𝐶1𝐵=(𝐶𝐵)𝐶1𝐵=𝐵𝐵𝐶1𝐵=𝐵(𝐼𝐶1𝐵)=𝐵𝐶1(𝐶𝐵)=𝐵𝐶1𝐴.(ii)(𝐴1+𝐵1)𝐷=𝐴1𝐷+𝐵1𝐷=𝐴1𝐴𝐶1𝐵+𝐵1𝐵𝐶1𝐴=𝐶1(𝐵+𝐴)=𝐶1𝐶=𝐼.

Proof of Proposition 3.1. (ii) Directly stems from Lemma 3.2, item (ii) by posing 𝐴=𝐡𝑚𝑛,𝑘(𝑙𝑛,𝑘,𝑚𝑛,𝑘), 𝐵=𝐡𝑚𝑛,𝑘(𝑚𝑛,𝑘,𝑟𝑛,𝑘), and 𝐶=𝐴+𝐵=𝐡𝑚𝑛,𝑘(𝑙𝑛,𝑘,𝑟𝑛,𝑘). Indeed, the lemma implies that𝐷1=𝐴1𝐶𝐵1=𝐡𝑚𝑛,𝑘𝑙𝑛,𝑘,𝑚𝑛,𝑘1𝐡𝑚𝑛,𝑘𝑙𝑛,𝑘,𝑟𝑛,𝑘𝐡𝑚𝑛,𝑘𝑙𝑛,𝑘,𝑚𝑛,𝑘1=𝚺1𝑛,𝑘.(3.4)(i)We have𝐋𝑛,𝑘+𝐑𝑛,𝑘=𝐠𝑚𝑛,𝑘𝑇𝐡𝑙𝑛,𝑘,𝑚𝑛,𝑘1+𝐡𝑚𝑛,𝑘,𝑟𝑛,𝑘1𝝈𝑛,𝑘=𝐠𝑚𝑛,𝑘𝑇𝚺1𝑛,𝑘𝝈𝑛,𝑘=𝐠𝑚𝑛,𝑘𝑇𝝈1𝑛,𝑘𝑇,(3.5) which ends the demonstration of the proposition.

Let us define 𝐋0,0=(𝐡(𝑙0,0,𝑟0,0))1𝐠1(𝑟0,0)𝝈0,0. With this notations we define the functions in a compact form as follows.

Definition 3.3. For every (𝑛,𝑘) in with 𝑛0, the continuous functions 𝝍𝑛,𝑘 are defined on their support 𝑆𝑛,𝑘 as 𝝍𝑛,𝑘(𝑡)=𝐠(𝑡)𝐡𝑙𝑛,𝑘,𝑡𝐋𝑛,𝑘,𝑙𝑛,𝑘𝑡𝑚𝑛,𝑘,𝐠(𝑡)𝐡𝑡,𝑟𝑛,𝑘𝐑𝑛,𝑘,𝑚𝑛,𝑘𝑡𝑟𝑛,𝑘,(3.6) and the basis element 𝝍0,0 is given on [0,1] by 𝝍0,0(𝑡)=𝐠(𝑡)𝐡𝑙0,0,𝑡𝐋0,0.(3.7)

The definition implies that 𝝍𝑛,𝑘 are continuous functions in the space of piecewise derivable functions with piecewise continuous derivative which takes value zero at zero. We denote such a space by 𝐶10([0,1],𝑑×𝑑).

Before studying the property of the functions 𝝍𝑛,𝑘, it is worth remembering that their definitions include the choice of a square root 𝝈𝑛,𝑘 of 𝚺𝑛,𝑘. Properly speaking, there is thus a class of bases 𝝍𝑛,𝑘 and all the points we develop in the sequel are valid for this class. However, for the sake of simplicity, we consider from now on that the basis under scrutiny results from choosing the unique square root 𝝈𝑛,𝑘 that is lower triangular with positive diagonal entries (the Cholesky decomposition).

3.1.1. Underlying System of Orthonormal Functions

We first introduce a family of functions 𝜙𝑛,𝑘 and show that it constitutes an orthogonal basis on a certain Hilbert space. The choice of this basis can seem arbitrary at first sight, but the definition of these function will appear natural for its relationship with the functions 𝝍𝑛,𝑘 and Φ𝑛,𝑘 that is made explicit in the sequel, and the mathematical rigor of the argument lead us to choose this apparently artificial introduction.

Definition 3.4. For every (𝑛,𝑘) in with 𝑛0, we define a continuous function 𝜙𝑛,𝑘[0,1]𝑚×𝑑 which is zero outside its support 𝑆𝑛,𝑘 and has the expressions: 𝜙𝑛,𝑘(𝑡)=𝐟(𝑡)𝑇𝐋𝑛,𝑘if𝑙𝑛,𝑘𝑡<𝑚𝑛,𝑘,𝐟(𝑡)𝑇𝐑𝑛,𝑘if𝑚𝑛,𝑘𝑡<𝑟𝑛,𝑘.(3.8) The basis element 𝜙0,0 is defined on [0,1] by 𝜙0,0(𝑡)=𝐟(𝑡)𝑇𝐋0,0.(3.9)

Remark that the definitions make apparent the fact that these two families of functions are linked for all (𝑛,𝑘) in through the simple relation𝝍𝑛,𝑘=𝜶𝝍𝑛,𝑘+𝚪𝜙𝑛,𝑘.(3.10) Moreover, this collection of functions 𝜙𝑛,𝑘 constitutes an orthogonal basis of functions, in the following sense.

Proposition 3.5. Let 𝐿2𝐟 be the closure of 𝐮[0,1]𝑚𝐯𝐿2[0,1],𝑑,𝐮=𝐟𝑇𝐯,(3.11) equipped with the natural norm of 𝐿2([0,1],𝑚). It is a Hilbert space, and moreover, for all 0𝑗<𝑑, the family of functions 𝑐𝑗(𝜙𝑛,𝑘) defined as the columns of 𝜙𝑛,𝑘, namely 𝑐𝑗𝜙𝑛,𝑘=𝜙𝑛,𝑘𝑖,𝑗0𝑖<𝑚,(3.12) forms a complete orthonormal basis of 𝐿2𝐟.

Proof. The space 𝐿2𝐟 is clearly a Hilbert space as a closed subspace of the larger Hilbert space 𝐿2([0,1],𝑚) is equipped with the standard scalar product: 𝐮,𝐯𝐿2[0,1],𝑑,(𝐮,𝐯)=10𝐮(𝑡)𝑇𝐯(𝑡)𝑑𝑡.(3.13) We now proceed to demonstrate that the columns of 𝜙𝑛,𝑘 form an orthonormal family which generates a dense subspace of 𝐿2𝐟. To this end, we define 𝑀([0,1],𝑚×𝑑) as the space of functions 𝐀[0,1]𝑚×𝑑𝑗0𝑗<𝑑,𝑡𝐀𝑖,𝑗(𝑡)0𝑖<𝑚𝐿2([0,1],𝑚),(3.14) that is, the space of functions that take values in the set of 𝑚×𝑑-matrices whose columns are in 𝐿2([0,1],𝑚). This definition allows us to define the bilinear function 𝒫𝑀([0,1],𝑚×𝑑)×𝑀([0,1],𝑚×𝑑)𝑑×𝑑 as 𝒫(𝐀,𝐁)=10𝐀(𝑡)𝑇𝐁(𝑡)𝑑𝑡satisfying𝒫(𝐁,𝐀)=𝒫(𝐀,𝐁)𝑇,(3.15) and we observe that the columns of 𝜙𝑛,𝑘 form an orthonormal system if and only if ((𝑝,𝑞),(𝑛,𝑘))×,𝒫𝜙𝑛,𝑘,𝜙𝑝,𝑞=10𝜙𝑛,𝑘(𝑡)𝑇𝜙𝑝,q(𝑡)𝑑𝑡=𝛿𝑛,𝑘𝑝,𝑞𝐈𝑑,(3.16) where 𝛿𝑛,𝑘𝑝,𝑞 is the Kronecker delta function, whose value is 1 if 𝑛=𝑝 and 𝑘=𝑞, and 0 otherwise.
First of all, since the functions 𝜙𝑛,𝑘 are zero outside the interval 𝑆𝑛,𝑘, the matrix 𝒫(𝜙𝑛,𝑘,𝜙𝑝,𝑞) is nonzero only if 𝑆𝑛,𝑘𝑆𝑝,𝑞. In such cases, assuming that 𝑛𝑝 and, for example, that 𝑛<𝑝, we necessarily have 𝑆𝑛,𝑘 strictly included in 𝑆𝑝,𝑞: more precisely, 𝑆𝑛,𝑘 is either included in the left-child support 𝑆𝑝+1,2𝑞 or in the right-child support 𝑆𝑝+1,2𝑞+1 of 𝑆𝑝,𝑞. In both cases, writing the matrix 𝒫(𝜙𝑛,𝑘(𝑡),𝜙𝑝,𝑞) shows that it is expressed as a matrix product whose factors include 𝒫(𝜙𝑛,𝑘,𝐟𝑇). We then show that 𝒫𝜙𝑛,𝑘,𝐟𝑇=10𝜙𝑛,𝑘(𝑡)𝑇𝐟(𝑡)𝑇=𝐋𝑇𝑛,𝑘𝑚𝑛,𝑘𝑙𝑛,𝑘𝐟(𝑢)𝐟𝑇(𝑢)𝑑𝑢𝐑𝑇𝑛,𝑘𝑟𝑛,𝑘𝑚𝑛,𝑘𝐟(𝑢)𝐟𝑇(𝑢)𝑑𝑢,=𝐋𝑇𝑛,𝑘𝐡𝑙𝑛,𝑘,𝑚𝑛,𝑘𝐑𝑇𝑛,𝑘𝐡𝑚𝑛,𝑘,𝑟𝑛,𝑘=𝝈𝑇𝑛,𝑘𝐠1𝑚𝑛,𝑘𝑇𝝈𝑇𝑛,𝑘𝐠1𝑚𝑛,𝑘𝑇,(3.17) which entails that 𝒫(𝜙𝑛,𝑘,𝐟𝑇)=0 if 𝑛<𝑝. If 𝑛>𝑝, we remark that 𝒫(𝜙𝑛,𝑘,𝜙𝑝,𝑞)=𝒫(𝜙𝑝,𝑞,𝜙𝑛,𝑘)𝑇, and we conclude that 𝒫(𝜙𝑛,𝑘,𝜙𝑝,𝑞)=𝟎 from the preceding case. For 𝑛=𝑝, we directly compute for 𝑛>0 the only nonzero term 𝒫𝜙𝑛,𝑘,𝜙𝑛,𝑘=𝐋𝑇𝑛,𝑘𝑚𝑛,𝑘𝑙𝑛,𝑘𝐟(𝑢)𝐟𝑇(𝑢)𝑑𝑢𝐋𝑛,𝑘+𝐑𝑇𝑛,𝑘𝑟𝑛,𝑘𝑚𝑛,𝑘𝐟(𝑢)𝐟𝑇(𝑢)𝑑𝑢𝐑𝑛,𝑘,=𝝈𝑇𝑛,𝑘𝐠1𝑚𝑛,𝑘𝑇𝐡𝑙𝑛,𝑘,𝑚𝑛,𝑘1𝐠1𝑚𝑛,𝑘𝝈𝑛,𝑘+𝝈𝑇𝑛,𝑘𝐠1𝑚𝑛,𝑘𝑇𝐡𝑚𝑛,𝑘,𝑟𝑛,𝑘1𝐠1𝑚𝑛,𝑘𝝈𝑛,𝑘.(3.18) Using the passage relationship between the symmetric functions 𝐡 and 𝐡𝑚𝑛,𝑘 given in (2.7), we can then write 𝒫𝜙𝑛,𝑘,𝜙𝑛,𝑘=𝝈𝑇𝑛,𝑘𝐡𝑚𝑛,𝑘𝑙𝑛,𝑘,𝑚𝑛,𝑘1𝝈𝑛,𝑘+𝝈𝑇𝑛,𝑘𝐡𝑚𝑛,𝑘𝑚𝑛,𝑘,𝑟𝑛,𝑘1𝝈𝑛,𝑘.(3.19) Proposition 3.1 implies that 𝐡𝑚𝑛,𝑘(𝑙𝑛,𝑘,𝑚𝑛,𝑘)1+𝐡𝑚𝑛,𝑘(𝑚𝑛,𝑘,𝑟𝑛,𝑘)1=𝚺1𝑛,𝑘=(𝝈1𝑛,𝑘)𝑇𝝈1𝑛,𝑘 which directly implies that 𝒫(𝜙𝑛,𝑘,𝜙𝑇𝑛,𝑘)=𝐈𝑑. For 𝑛=0, a computation of the exact same flavor yields that 𝒫(𝜙0,0,𝜙0,0)=𝐈𝑑. Hence, we have proved that the collection of columns of 𝜙𝑛,𝑘 forms an orthonormal family of functions in 𝐿2𝐟 (the definition of 𝜙𝑛,𝑘 clearly states that its columns can be written in the form of elements of 𝐿2𝐟).
The proof now amounts showing the density of the family of functions we consider. Before showing this density property, we introduce for all (𝑛,𝑘) in the functions 𝐏𝑛,𝑘[0,1]𝑑×𝑑 with support on 𝑆𝑛,𝑘 defined by 𝐏𝑛,𝑘(𝑡)=𝐋𝑛,𝑘if𝑙𝑛,𝑘𝑡<𝑚𝑛,𝑘,𝐑𝑛,𝑘if𝑚𝑛,𝑘𝑡<𝑟𝑛,𝑘,𝑛0,𝐏0,0(𝑡)=𝐋0,0.(3.20) Showing that the family of columns of 𝜙𝑛,𝑘 is dense in 𝐿2𝐟 is equivalent to show that the column vectors of the matrices 𝐏𝑛,𝑘 seen as a function of 𝑡 are dense in 𝐿2([0,1],𝑑). It is enough to show that the span of such functions contains the family of piecewise continuous 𝑑-valued functions that are to be constant on 𝑆𝑛,𝑘, (𝑛,𝑘) in (the density of the endpoints of the partition 𝑁𝐷𝑁 entails that the latter family generates 𝐿2([0,1],𝑑)).
In fact, we show that the span of functions 𝑉𝑁=span𝑡𝑐𝑗𝐏𝑛,𝑘(𝑡)0𝑗<𝑑,(𝑛,𝑘)𝑁(3.21) is exactly equal to the space 𝐾𝑁 of piecewise continuous functions from [0,1] to 𝑑 that are constant on the supports 𝑆𝑁+1,𝑘, for any (𝑁+1,𝑘) in . The fact that 𝑉𝑁 is included in 𝐾𝑁 is clear from the fact that the matrix-valued functions 𝐏𝑁,𝑘 are defined constant on the support 𝑆𝑁+1,𝑘, for (𝑁,𝑘) in 𝐼.
We prove that 𝐾𝑁 is included in 𝑉𝑁 by induction on 𝑁0. The property is clearly true at rank 𝑁=0 since 𝐏0,0 is then equal to the constant invertible matrix 𝐋0,0. Assuming that the proposition true at rank 𝑁1 for a given 𝑁>0, let us consider a piecewise continuous function 𝐜[0,1]𝑑 in 𝐾𝑁1. Remark that, for every (𝑁,𝑘) in , the function 𝐜 can only take two values on 𝑆𝑁,𝑘 and can have discontinuity jump in 𝑚𝑁,𝑘: let us denote these jumps as 𝐝𝑁,𝑘=𝐜𝑚+𝑁,𝑘𝐜𝑚𝑁,𝑘.(3.22) Now, remark that for every (𝑁,𝑘) in , the matrix-valued functions 𝐏𝑁,𝑘 take only two matrix values on 𝑆𝑁,𝑘, namely, 𝐋𝑁,𝑘 and 𝐑𝑁,𝑘. From Proposition 3.1, we know that 𝐋𝑁,𝑘+𝐑𝑁,𝑘=𝐌𝑁,𝑘 is invertible. This fact directly entails that there exist vectors 𝐚𝑁,𝑘, for any (𝑁,𝑘) in , such that 𝐝𝑁,𝑘=(𝐋𝑁,𝑘+𝐑𝑁,𝑘)(𝐚𝑁,𝑘). We then necessarily have that the function 𝐜=𝐜+𝐏𝑛,𝑘𝐚𝑛,𝑘 is piecewise constant on the supports 𝑆𝑁,𝑘, (𝑁,𝑘) in . By recurrence hypothesis, 𝐜 belongs to 𝑉𝑁1, so that 𝐜 belongs to 𝑉𝑁, and we have proved that 𝐾𝑁𝑉𝑁. Therefore, the space generated by the column vectors 𝑃𝑛,𝑘 is dense in 𝐿2[0,1], which completes the proof that the functions 𝑡[(𝜙𝑛,𝑘(𝑡))𝑖,𝑗]0𝑖<𝑚 form a complete orthonormal family of 𝐿2[0,1].

The fact that the column functions of 𝜙𝑛,𝑘 form a complete orthonormal system of 𝐿2𝐟 directly entails the following decomposition of the identity on 𝐿2𝐟.

Corollary 3.6. If 𝛿 is the real delta Dirac function, one has (𝑛,𝑘)𝜙𝑛,𝑘(𝑡)𝜙𝑇𝑛,𝑘(𝑠)=𝛿(𝑡𝑠)𝐼𝑑𝐿2𝐟.(3.23)

Proof. Indeed it easy to verify that, for all 𝐯 in 𝐿2𝐟, we have for all 𝑁>0𝑈(𝑛,𝑘)𝑁𝜙𝑛,𝑘(𝑡)𝜙𝑇𝑛,𝑘(𝑠)𝐯(𝑠)𝑑𝑠=(𝑛,𝑘)𝑁𝜙𝑛,𝑘(𝑡)𝒫𝜙𝑛,𝑘,𝐯=(𝑛,𝑘)𝑁𝑑1𝑝=0𝑐𝑝𝜙𝑛,𝑘𝑐𝑝𝜙𝑛,𝑘,𝐯,(3.24) where (𝑐𝑝(𝜙𝑛,𝑘),𝐯) denotes the inner product in 𝐿2𝐟 between 𝐯 and the 𝑝-column of 𝝍𝑛,𝑘. Therefore, by the Parseval identity, we have in the 𝐿2𝐟 sense 𝑈(𝑛,𝑘)𝜙𝑛,𝑘(𝑡)𝜙𝑇𝑛,𝑘(𝑠)𝐯(𝑠)𝑑𝑠=𝐯(𝑡).(3.25)

From now on, abusing language, we will say that the family of 𝑚×𝑑-valued functions 𝜙𝑛,𝑘 is an orthonormal family of functions to refer to the fact that the columns of such matrices form orthonormal set of 𝐿2𝐟. We now make explicit the relationship between this orthonormal basis and our functions (𝝍𝑛,𝑘) derived in our analysis of the multidimensional Gauss-Markov processes.

3.1.2. Generalized Dual Operators

The Integral Operator 𝓚
The basis 𝜙𝑛,𝑘 is of great interest in this paper for its relationship to the functions 𝝍𝑛,𝑘 that naturally arise in the decomposition of the Gauss-Markov processes. Indeed, the collection 𝝍𝑛,𝑘 can be generated from the orthonormal basis 𝜙𝑛,𝑘 through the action of the integral operator 𝒦 defined on 𝐿2([0,1],𝑚) into 𝐿2([0,1],𝑑) by 𝐮𝒦[𝐮]=𝑡𝐠(𝑡)𝑈𝟙[0,𝑡](𝑠)𝐟(𝑠)𝐮(𝑠)𝑑𝑠,(3.26) where 𝑈[0,1] is an open set and, for any set 𝐸𝑈,𝟙𝐸() denotes the indicator function of 𝐸. Indeed, realizing that 𝒦 acts on 𝑀([0,1],𝑚×𝑑) into 𝑀([0,1],𝑑×𝑑) through 𝐀𝑀[0,1],m×𝑑,𝒦[𝐀]=𝒦𝑐0(𝐀),,𝒦𝑐𝑑1(𝐀),(3.27) where 𝑐𝑗(𝐀) denotes the 𝑗th 𝑚-valued column function of 𝐀, we easily see that for all (𝑛,𝑘) in , 0𝑡1, 𝝍𝑛,𝑘(𝑡)=𝐠(𝑡)𝑡0𝐟(𝑠)𝜙𝑛,𝑘(𝑠)𝑑𝑠=𝒦𝜙𝑛,𝑘(𝑡).(3.28) It is worth noticing that the introduction of the operator 𝒦 can be considered natural since it characterizes the centered Gauss-Markov process 𝐗 through loosely writing 𝐗=𝒦[𝐝𝐖].
In order to exhibit a dual family of functions to the basis 𝝍𝑛,𝑘, we further investigate the property of the integral operator 𝒦. In particular, we study the existence of an inverse operator 𝒟, whose action on the orthonormal basis 𝜙𝑛,𝑘 will conveniently provide us with a dual basis to 𝝍𝑛,𝑘. Such an operator does not always exist; nevertheless, under special assumptions, it can be straightforwardly expressed as a generalized differential operator.

The Differential Operator 𝓓
Here, we make the assumptions that 𝑚=𝑑, that, for all 𝑡, 𝐟(𝑡) is invertible in 𝑑×𝑑, and that 𝐟 and 𝐟1 have continuous derivatives, which especially implies that 𝐿2𝐟=𝐿2(𝑑). In this setting, we define the space 𝐷0(𝑈,𝑑) of functions in 𝐶0(𝑈,𝑑) that are zero at zero and denote by 𝐷0(𝑈,𝑑) its dual in the space of distributions (or generalized functions). Under the assumptions just made, the operator 𝒦𝐷0(𝑈,𝑑)𝐷0(𝑈,𝑑) admits the differential operator 𝒟𝐷0(𝑈,𝑑)𝐷0(𝑈,𝑑) defined by 𝐮𝐷0𝑈,𝑑𝒟[𝐮]=𝑡𝐟1(𝑡)𝑑𝑑𝑡𝐠1(𝑡)𝐮(𝑡)(3.29) as its inverse, that is, when restricted to 𝐷0(𝑈,𝑑), we have 𝒟𝒦=𝒦𝒟=𝐼𝑑 on 𝐷0(𝑈,𝑑). The dual operators of 𝒦 and 𝒟 are expressed, for any 𝐮 in 𝐷0(𝑈,𝑑), as 𝒟[𝐮]=𝑡𝐠1(𝑡)𝑇𝑑𝑑𝑡𝐟1(𝑡)𝑇𝐮(𝑡),𝒦[𝐮]=𝑡𝐟(𝑡)𝑇𝑈𝟙[0,𝑡](𝑠)𝐠𝑇(𝑠)𝐮(𝑠)𝑑𝑠.(3.30) They satisfy (from the properties of 𝒦 and 𝒟) 𝒟𝒦=𝒦𝒟=𝐼𝑑 on 𝐷0(𝑈,𝑑). By dual pairing, we extend the definition of the operators 𝒦, 𝒟 as well as their dual operators, to the space of generalized function 𝐷0(𝑈,𝑑). In details, for any distribution 𝑇 in 𝐷0(𝑈,𝑑) and test function 𝐮 in 𝐷0(𝑈,𝑑), define 𝒦 and 𝒦 by (𝒟[𝑇],𝐮)=𝑇,𝒟[𝐮],(𝒦[𝑇],𝐮)=𝑇,𝒦[𝐮],(3.31) and reciprocally for the dual operators 𝒟 and 𝒦.

Candidate Dual Basis
We are now in a position to use the orthonormality of 𝜙𝑛,𝑘 to infer a dual family of the basis 𝝍𝑛,𝑘. For any function 𝐮 in 𝐿2(𝑈,𝑑), the generalized function 𝒦[𝐮] belongs to 𝐶0(𝑈,𝑑), the space of continuous functions that are zero at zero. We equip this space with the uniform norm and denote its topological dual 𝑅0(𝑈,𝑑), the set of 𝑑-dimensional Radon measures with 𝑅0(𝑈,𝑑)𝐷0(𝑈,𝑑). Consequently, operating in the Gelfand triple 𝐶0𝑈,𝑑𝐿2𝑈,𝑑𝑅0𝑈,𝑑,(3.32) we can write, for any function 𝐮, 𝐯 in 𝐿2(𝑈,𝑑)𝑅0(𝑈,𝑑), (𝐮,𝐯)=((𝒟𝒦)[𝐮],𝐯)=𝒦[𝐮],𝒟[𝐯].(3.33) The first equality stems from the fact that, when 𝒦 and 𝒟 are seen as generalized functions, they are still inverse of each other, so that in particular 𝒟𝒦=𝐼𝑑 on 𝐿2(𝑈,𝑑). The dual pairing associated with the Gelfand triple (3.32) entails the second equality where 𝒟 is the generalized operator defined on 𝐷0(𝑈,𝑑) and where 𝒟[𝐯] is in 𝑅0(𝑈,𝑑).
As a consequence, defining the functions 𝜹𝑛,𝑘 in 𝑅0(𝑈,𝑑×𝑑), the 𝑑×𝑑-dimensional space of Radon measures, by 𝜹𝑛,𝑘=𝒟𝜙𝑛,𝑘=𝒟𝑐𝑖𝜙𝑛,𝑘,,𝒟𝑐𝑗𝜙𝑛,𝑘(3.34) provides us with a family of 𝑑×𝑑-generalized functions which are dual to the family 𝝍𝑛,𝑘 in the sense that, for all ((𝑛,𝑘),(𝑝,𝑞)) in ×, we have 𝒫𝜹𝑛,𝑘,𝝍𝑝,𝑞=𝛿𝑛,𝑘𝑝,𝑞𝐈𝑑,(3.35) where the definition of 𝒫 has been extended through dual pairing: given any 𝐀 in 𝑅0(𝑈,𝑚×𝑑) and any 𝐁 in 𝐶0(𝑈,𝑚×𝑑), we have 𝒫(𝐀,𝐁)=𝑐𝑖(𝐀),𝑐𝑗(𝐁)0𝑖,𝑗<𝑑(3.36) with (𝑐𝑖(𝐀),𝑐𝑗(𝐁)) denoting the dual pairing between the 𝑖th column of 𝐀 taking value in 𝑅0(𝑈,𝑑) and the 𝑗th column of 𝐁 taking value in 𝐶0(𝑈,𝑑). Under the favorable hypothesis of this section, the 𝑑×𝑑-generalized functions 𝜹𝑛,𝑘 can actually be easily computed since considering the definition of 𝜙𝑛,𝑘 shows that the functions (𝐟1)𝑇𝜙𝑛,𝑘 have support 𝑆𝑛,𝑘 and are constant on 𝑆𝑛+1,2𝑘 and S𝑛+1,2𝑘+1 in 𝑑×𝑑. Only the discontinuous jumps in 𝑙𝑛,𝑘, 𝑚𝑛,𝑘, and 𝑟𝑛,𝑘 intervene, leading to expressing for (𝑛,𝑘) in , 𝑛0𝜹𝑛,𝑘(𝑡)=𝐠(𝑡)1𝑇𝐌𝑛,𝑘𝛿𝑡𝑚𝑛,𝑘𝐋𝑛,𝑘𝛿𝑡𝑙𝑛,𝑘+𝐑𝑛,𝑘𝛿𝑡𝑟𝑛,𝑘(3.37) and 𝜹0,0(𝑡)=(𝐠(𝑡)1)𝑇𝐋0,0, where 𝛿() denotes the standard delta Dirac function (centered in 0). These functions can be extended to the general setting of the paper since its expressions do not involve the assumptions made on the invertibility and smoothness of 𝐟(𝑡). We now show that these functions, when defined in the general setting, still provide a dual basis of the functions 𝝍𝑛,𝑘.

3.1.3. Dual Basis of Generalized Functions

The expression of the basis 𝜹𝑛,𝑘 that has been found under favorable assumptions makes no explicit reference to these assumptions. It suggests defining functions 𝜹𝑛,𝑘 formally as linear combination of Dirac functions acting by duality on 𝐶0(𝑈,𝑑×𝑑).

Definition 3.7. For (𝑛,𝑘) in , the family of generalized functions 𝜹𝑛,𝑘 in 𝑅0(𝑈,𝑑×𝑑) is given by (𝑛0) 𝜹𝑛,𝑘(𝑡)=𝐠(𝑡)1𝑇𝐌𝑛,𝑘𝛿𝑡𝑚𝑛,𝑘𝐋𝑛,𝑘𝛿𝑡𝑙𝑛,𝑘+𝐑𝑛,𝑘𝛿𝑡𝑟𝑛,𝑘,(3.38) and 𝜹0,0(𝑡)=(𝐠(𝑡)1)𝑇𝐋0,0, where 𝛿 is the standard Dirac distribution.

Notice that the basis 𝜹𝑛,𝑘 is defined for the open set 𝑈. For the sake of consistency, we extend the definition of the families 𝝍𝑛,𝑘 and 𝜙𝑛,𝑘 on 𝑈 by setting them to zero on 𝑈[0,1], except for 𝝍0,0, which is continued for 𝑡>1 by a continuous function 𝐜 that is compactly supported in [1,𝑎) for a given 𝑎 in 𝑈, 𝑎>1 and satisfies 𝐜(1)=𝝍0,0(1).

We can now formulate the following.

Proposition 3.8. Given the dual pairing in 𝐶0(𝑈)𝐿2(𝑈)𝑅(𝑈) where 𝑈 is a bounded open set of containing [0,1], the family of continuous functions 𝝍𝑛,𝑘 in 𝐶0(𝑈) admits, for dual family in 𝑅(𝑈), the set of distributions 𝜹𝑛,𝑘.

Proof. We have to demonstrate that, for all ((𝑛,𝑘),(𝑝,𝑞)) in ×, 𝒫𝜹𝑝,𝑞,𝝍𝑛,𝑘=𝛿𝑛,𝑘𝑝,𝑞𝐈𝑑.(3.39) Suppose first that 𝑛,𝑝>0. If 𝑝<𝑛, 𝒫(𝜹𝑛,𝑘,𝝍𝑝,𝑞) can only be nonzero if the support 𝑆𝑝,𝑞 is strictly included in 𝑆𝑛,𝑘. We then have 𝒫𝜹𝑝,𝑞,𝝍𝑛,𝑘=𝐌𝑇𝑝,𝑞𝐠1𝑚𝑝,𝑞𝝍𝑛,𝑘𝑚𝑝,𝑞𝐋𝑇𝑝,𝑞𝐠1𝑙𝑝,𝑞𝝍𝑛,𝑘𝑙𝑝,𝑞+𝐑𝑇𝑝,𝑞𝐠1𝑟𝑝,𝑞𝝍𝑛,𝑘𝑟𝑝,𝑞.(3.40) Assume that 𝑆𝑝,𝑞 is to the left of 𝑚𝑛,𝑘, that is, 𝑆𝑝,𝑞 is a left child of 𝑆𝑛,𝑘 in the nested binary tree of supports and write 𝒫𝜹𝑝,𝑞,𝝍𝑛,𝑘=𝐌𝑇𝑝,𝑞𝐡𝑙𝑛,𝑘,𝑚𝑝,𝑞𝐋𝑇𝑝,𝑞𝐡𝑙𝑛,𝑘,𝑙𝑝,𝑞𝐑𝑇𝑝,𝑞𝐡𝑙𝑛,𝑘,𝑟𝑝,𝑞𝐋𝑛,𝑘.(3.41) Using the fact that 𝐌𝑝,𝑞=𝐋𝑝,𝑞+𝐑𝑝,𝑞 and that the function (𝑥,𝑦), as any integral between 𝑥 and 𝑦, satisfies the chain rule 𝐡(𝑥,𝑦)=𝐡(𝑥,𝑧)+𝐡(𝑧,𝑦) for all (𝑥,𝑦,𝑧), we obtain 𝒫𝜹𝑝,𝑞,𝝍𝑛,𝑘=𝐋𝑇𝑝,𝑞𝐡𝑙𝑛,𝑘,𝑚𝑝,𝑞𝐡𝑙𝑛,𝑘,𝑙𝑝,𝑞+𝐑𝑇𝑝,𝑞𝑙𝑛,𝑘,𝑚𝑝,𝑞𝑙𝑛,𝑘,𝑟𝑝,𝑞𝐋𝑛,𝑘=𝐋𝑇𝑝,𝑞𝐡𝑙𝑝,𝑞,𝑚𝑝,𝑞+𝐑𝑇𝑝,𝑞𝑟𝑝,𝑞,𝑚𝑝,𝑞𝐋𝑛,𝑘=𝝈𝑇𝑝,𝑞𝐠1𝑚𝑝,𝑞𝑇𝐡𝑙𝑝,𝑞,𝑚𝑝,𝑞1𝑇𝐡𝑙𝑝,𝑞,𝑚𝑝,𝑞+𝝈𝑇𝑝,𝑞𝐠1𝑚𝑝,𝑞𝑇𝐡𝑚𝑝,𝑞,𝑟𝑝,𝑞1𝑇𝐡𝑚𝑝,𝑞,𝑟𝑝,𝑞𝐋𝑛,𝑘=𝟎.(3.42) The same result is true if 𝑆𝑝,𝑞 is a right child of 𝑆𝑛,𝑘 in the nested binary tree of supports. If 𝑝=𝑛, necessarily the only nonzero term is for 𝑞=𝑝, that is, 𝒫𝜹𝑝,𝑞,𝝍𝑛,𝑘=𝐌𝑇𝑛,𝑘𝐠1𝑚𝑛,𝑘𝝍𝑚𝑛,𝑘=𝐌𝑇𝑛,𝑘𝐡𝑙𝑛,𝑘,𝑚𝑛,𝑘𝐋𝑛,𝑘=𝝈1𝑝,𝑞𝐠𝑚𝑛,𝑘𝐡l𝑛,𝑘,𝑚𝑛,𝑘𝐡𝑙𝑛,𝑘,𝑚𝑛,𝑘1𝐠1𝑚𝑛,𝑘𝝈𝑝,𝑞=𝐈𝑑.(3.43) If 𝑝>𝑛, 𝒫(𝜹𝑛,𝑘,𝝍𝑝,𝑞) can only be nonzero if the support 𝑆𝑛,𝑘 is included in 𝑆𝑝,𝑞, but then 𝝍𝑛,𝑘 is zero in 𝑙𝑝,𝑞, 𝑚𝑝,𝑞, 𝑟𝑝,𝑞 so that 𝒫(𝜹𝑛,𝑘,𝝍𝑝,𝑞)=𝟎.
Otherwise, if 𝑛=0 and 𝑝>0, we directly have 𝒫𝜹𝑝,𝑞,𝝍0,0=𝐌𝑇𝑝,𝑞𝐠1𝑚𝑝,𝑞𝝍0,0𝑚𝑝,𝑞𝐋𝑇𝑝,𝑞𝐠1𝑙𝑝,𝑞𝝍0,0𝑙𝑝,𝑞+𝐑𝑇𝑝,𝑞𝐠1𝑟𝑝,𝑞𝝍0,0𝑟𝑝,𝑞,=𝐋𝑇𝑝,𝑞𝐡𝑙𝑝,𝑞,𝑚𝑝,𝑞+𝐑𝑇𝑝,𝑞𝐡𝑚𝑝,𝑞,𝑟𝑝,𝑞𝐋0,0,=𝝈𝑇𝑝,𝑞𝐠1𝑚𝑝,𝑞𝑇+𝝈𝑇𝑝,𝑞𝐠1𝑚𝑝,𝑞𝑇𝐋𝑛,𝑘,=𝟎.(3.44) Finally, if 𝑝=0, given the simple form of 𝜹0,0 with a single Dirac function centered in 𝑟0,0, we clearly have 𝒫(𝜹0,0,𝝍𝑛,𝑘)=0, and if 𝑛>0𝒫𝜹0,0,𝝍0,0,=𝐋𝑇0,0𝐡𝑙0,0,𝑟0,0𝐋0,0,=𝝈𝑇0,0𝐠1𝑟0,0𝑇𝐋0,0,=𝝈𝑇0,0𝐠1𝑟0,0𝑇𝐡𝑙0,0,𝑟0,0𝐋0,01𝐠1𝑟0,0𝝈0,0,=𝝈𝑇0,0𝐂1𝑟0,0𝝈0,0,(3.45) and using the fact that (by definition of 𝝈0,0) we have 𝝈0,0𝝈𝑇0,0=𝐂𝑟0,0, this last expression is equal to 𝒫𝜹0,0,𝝍0,0=𝝈𝑇0,0𝝈𝑇0,01𝝈0,01𝝈0,0=𝐼𝑑,(3.46) which completes the proof.

This proposition directly implies the main result of the section.

Theorem 3.9. The collection of functions (𝝍𝑛,𝑘;(𝑛,𝑘)𝐼) constitutes a Schauder basis of functions on 𝐶0(𝑈,𝑑), that is, any element of 𝐶0(𝑈,𝑑) can be written in a unique way as a sum of coefficients 𝑎𝑛,𝑘 multiplied by 𝝍𝑛,𝑘.

This theorem provides us with a complementary view of stochastic processes: in addition to the standard sample paths view, this structure allows to see the Gauss-Markov processes as coefficients on the computed basis. This duality is developed in the sequel.

3.2. The Sample Paths Space
3.2.1. The Construction Application

The Schauder basis of functions with compact supports constructed allows to define functions by considering the coefficients on this basis, which constitute sequences of real numbers in the space𝜉Ω=𝝃=𝝃𝑛,𝑘𝐼;(𝑛,𝑘),𝝃𝑛,𝑘𝑑=𝑑.(3.47) We equip 𝜉Ω with the uniform norm 𝝃=sup(𝑛,𝑘)|𝝃𝑛,𝑘|, where we write |𝝃𝑛,𝑘|=sup0𝑖<𝑑|(𝝃𝑛,𝑘)𝑖|. We denote by (𝜉Ω) the Borelian sets of the topology induced by the uniform norm and we recall that 𝐶(𝜉Ω), the cylinder sets of 𝜉Ω, form a generative family of Borelian sets. Remark that not any sequence of coefficients provides a continuous function, and one needs to assume a certain decrease in the coefficients to get convergence. A sufficient condition to obtain convergent sequences is to consider coefficients in the space𝜉Ω=𝝃𝜉Ω𝛿(0,1),𝑁,(𝑛,𝑘)𝑁,||𝝃𝑛,𝑘||<2𝑛𝛿/2.(3.48) This set is clearly a Borelian set of 𝜉Ω since it can be written as a countable intersection and union of cylinder, namely, by denoting by 𝒥 the set of finite subset of and 𝛿𝑝=11/𝑝, 𝑝>1,𝜉Ω=𝑝>1𝐽𝒥𝑛𝐽𝝃𝜉Ωmax0𝑘<2𝑛1||𝝃𝑛,𝑘||<2𝑛𝛿𝑝/2.(3.49) It is also easy to verify that it forms a vectorial subspace of 𝜉Ω.

After these definitions, we are in position to introduce the following useful function.

Definition 3.10. One denotes by Ψ𝑁 the partial construction application: 𝚿𝑁=𝜉Ω,𝐶0[0,1],𝑑,𝝃,(𝑛,𝑘)𝑁𝝍𝑛,𝑘(𝑡)𝝃𝑛,𝑘,(3.50) where the 𝐶0([0,1],𝑑) is the 𝑑-dimensional Wiener space, which is complete under the uniform norm 𝐱=sup0𝑡1|𝐱(𝑡)|.

This sequence of partial construction applications is shown to converge to the construction application in the following.

Proposition 3.11. For every 𝝃 in 𝜉Ω, Ψ𝑁(𝝃) converges uniformly toward a continuous function in 𝐶0([0,1],𝑑). One will denote this function Ψ(𝝃), defined as 𝚿𝜉Ω𝐶0[0,1],𝑑,𝝃(𝑛,𝑘)𝜓𝑛,𝑘(𝑡)𝝃𝑛,𝑘,(3.51) and this application will be referred to as the construction application.

This proposition is proved in Appendix D. The image of this function constitutes a subset of the Wiener space continuous functions 𝐶0([0,1],𝑑). Let us now define the vectorial subspace 𝑥Ω=Ψ(𝜉Ω) of 𝐶0([0,1],𝑑) so that Ψ appears as a bijection.

It is important to realize that, in the multidimensional case, the space 𝑥Ω depends on Γ and 𝜶 in a nontrivial way. For instance, assuming that 𝜶=0, the space 𝑥Ω depends obviously crucially on the rank of Γ. To fix the idea, for a given constant Γ(𝐭)=[0,01]𝑇 in 𝑑×1, we expect the space 𝑥Ω to only include sample paths of 𝐶0([0,1],𝑑) for which the 𝑛1 first components are constant. Obviously, a process with such sample paths is degenerated in the sense that its covariance matrix is not invertible.

Yet, if we additionally relax the hypothesis that 𝜶0, the space 𝑥Ω can be dramatically altered: if we take𝜶(𝑡)=0110,(3.52) the space 𝑥Ω will represent the sample space of the 𝑑1-integrated Wiener process, a nondegenerate 𝑑-dimensional process we fully develop in the example section.

However, the situation is much simpler in the one-dimensional case: because the uniform convergence of the sample paths is preserved as long as 𝛼 is continuous and Γ is nonzero through (D.8), the definition 𝑥Ω does not depend on 𝛼 or Γ. Moreover, in this case, the space 𝑥Ω is large enough to contain reasonably regular functions as proved in Appendix D, Proposition 3.

In the case of the 𝑑1-integrated Wiener process, the space 𝑥Ω clearly contains the functions {𝐟=(𝑓𝑑1,,𝑓0)𝑓0𝐻,𝑓𝑖=𝑓𝑖1,0<𝑖<𝑑}.

This remark does not hold that the space 𝑥Ω does not depend on 𝛼 as long as 𝛼 is continuous because the uniform convergence of the sample paths is preserved through the change of basis of expansion 𝜓𝑛,𝑘 through (D.8).

We equip the space 𝑥Ω with the topology induced by the uniform norm on 𝐶0([0,1],𝑑). As usual, we denote (𝑥Ω) the corresponding Borelian sets. We prove in Appendix D the following.

Proposition 3.12. The function Ψ(𝜉Ω,(𝜉Ω))(𝑥Ω,(𝑥Ω)) is a bounded continuous bijection.

We therefore conclude that we dispose of a continuous bijection mapping the coefficients onto the sample paths, Ψ. We now turn to study its inverse, the coefficient application, mapping sample paths on coefficients over the Schauder basis.

3.2.2. The Coefficient Application

In this section, we introduce and study the properties of the following function.

Definition 3.13. One calls coefficient application and denotes by Ξ the function defined by 𝚵𝐶0[0,1],𝑑𝜉Ω=𝑑,𝐱𝚫(𝐱)={𝚫(𝐱)}(𝑛,𝑘)with{𝚫(𝐱)}𝑛,𝑘=𝒫𝜹𝑛,𝑘,𝐱.(3.53)

Should a function 𝑥 admit a uniformly convergent decomposition in terms of the basis of elements 𝝍𝑛,𝑘, the function Δ gives its coefficients in such a representation. More precisely, we have the following.

Theorem 3.14. The function Δ(𝑥Ω,(𝑥Ω))(𝜉Ω,(𝜉Ω)) is a measurable linear bijection whose inverse is Ψ=Δ1.

The proof of this theorem is provided in Appendix D.

4. Representation of Gauss-Markov Processes

4.1. Inductive Construction of Gauss-Markov Processes

Up to this point, we have rigorously defined the dual spaces of sample paths 𝑥Ω and coefficients 𝜉Ω. Through the use of the Schauder basis 𝝍𝑛,𝑘 and its dual family of generalized functions 𝜹𝑛,𝑘, we have defined the inverse measurable bijections Ψ and Δ transforming one space into the other. In doing so, we have unraveled the fundamental role played by the underlying orthonormal basis 𝜙𝑛,𝑘. We now turn to use this framework to formulate a pathwise construction of the Gauss-Markov processes in the exact same flavor as the Lévy-Ciesielski construction of the Wiener process.

4.1.1. Finite-Dimensional Approximations

Considering the infinite-dimensional subspace 𝑥Ω of 𝐶0([0,1],𝑑), let us introduce the equivalence relation 𝑁 as𝐱𝑁𝐲𝑡𝐷𝑁,𝐱(𝑡)=𝐲(𝑡).(4.1) We can use the functions Ψ to carry through the structure of 𝑁 on the infinite-dimensional space of coefficients 𝜉Ω:𝝃𝑁𝜼𝚿(𝜉)𝑁𝚿(𝜼)(𝑛,𝑘)𝑁,𝝃𝑛,𝑘=𝜼𝑛,𝑘,(4.2) which clearly entails that 𝐱𝑁𝐲 if and only if Δ(𝐱)𝑁Δ(𝐲). We denote the sets of equivalence classes of 𝑥Ω/𝑁=𝑥Ω𝑁 and 𝜉Ω/𝑁=𝜉Ω𝑁, which are both clearly isomorphic 𝑥Ω𝑁=(𝑑)=𝜉Ω𝑁. For every 𝑁>0, we define the finite-dimensional operators Ψ𝑁=𝑥𝐢𝑁Ψ𝜉𝐩𝑁 and Δ𝑁=𝜉𝐢𝑁Δ𝑥𝐩𝑁, with the help of the canonical projections 𝜉𝐩𝑁𝜉Ω𝜉Ω𝑁, 𝑥𝐩𝑁𝑥Ω𝑥Ω𝑁 and the inclusion map 𝜉𝐢𝑁𝜉Ω𝑁𝜉Ω, 𝑥𝐢𝑁𝑥Ω𝑁𝑥Ω.

The results of the preceding sections straightforwardly extend on the equivalence classes, and in particular we see that the functions Ψ𝑁𝜉Ω𝑁𝑥Ω𝑁 and Δ𝑁𝑥Ω𝑁𝜉Ω𝑁 are linear finite-dimensional bijections satisfying Ψ𝑁=Δ𝑁1. We write 𝐞={𝐞𝑝,𝑞}(𝑝,𝑞) (resp., 𝐟={𝐟𝑝,𝑞}(𝑝,𝑞)), the canonical basis of 𝜉Ω𝑁 (resp., 𝑥Ω𝑁) when listed in the recursive dyadic order. In these bases, the matrices Ψ𝑁 and Δ𝑁 are lower block triangular. Indeed, denoting Ψ𝑁 in the natural bases 𝐞={𝐞𝑝,𝑞}(𝑝,𝑞) and 𝐟={𝐟𝑝,𝑞}(𝑝,𝑞) by𝚿𝑁=𝝍𝑛,𝑘𝑚𝑖,𝑗=𝚿𝑖,𝑗𝑛,𝑘,(4.3) where Ψ𝑖,𝑗𝑛,𝑘 is a 𝑑×𝑑 matrix, the structure of the nested support 𝑆𝑛,𝑘 entails the block-triangular structure (where only possibly nonzero coefficients are written):𝛼Ψ𝑁=𝝍0,00,0𝝍1,00,0𝝍1,01,0𝝍2,00,0𝝍2,01,0𝝍2,02,0𝝍2,10,0𝝍2,11,0𝝍2,12,1𝝍3,00,0𝝍3,01,0𝝍3,02,0𝝍3,03,0𝝍3,10,0𝝍3,11,0𝝍3,12,0𝝍3,13,1𝝍3,20,0𝝍3,21,0𝝍3,22,1𝝍3,23,2𝝍3,30,0𝝍3,31,0𝝍3,32,1𝝍3,33,3.(4.4)

Similarly, for the matrix representation of Δ𝑁 in the natural bases 𝑒𝑛,𝑘 and 𝑓𝑖,𝑗𝚫𝑁=𝚫𝑛,𝑘𝑖,𝑗(4.5) proves to have the following triangular form:𝚫𝑁=𝐠1𝑡0,0𝑇𝐌0,0𝐠1𝑡0,0𝑇𝐑1,0𝐠1𝑡1,0𝑇𝐌1,0𝐠1𝑡0,0𝑇𝐑2,0𝐠1𝑡1,0𝑇𝐌2,0𝐠1𝑡0,0𝑇𝐑2,1𝐠1𝑡1,0𝑇𝐋2,1𝐠1𝑡2,1𝑇𝐌2,1.(4.6) The duality property, Proposition 3.8, simply reads for all 0𝑛<𝑁 and 0𝑘<2𝑛1, 0𝑝<𝑁 and 0𝑝<2𝑞1𝒫𝜹𝑝,𝑞,𝝍𝑛,𝑘=(𝑛,𝑘)𝑁𝚫𝑝,𝑞𝑖,𝑗𝚿𝑖,𝑗𝑛,𝑘=𝛿𝑝,𝑞𝑛,𝑘𝐈𝑑,(4.7) that is, Δ𝑁Ψ𝑁=𝐼𝑑𝜉Ω𝑁. But because we are now in a finite-dimensional setting, we also have Ψ𝑁Δ𝑁=𝐼𝑑𝑥Ω𝑁:𝛿𝑖,𝑗𝑘,𝑙𝐈𝑑=(𝑝,𝑞)𝑁𝚿𝑖,𝑗𝑝,𝑞𝚫𝑝,𝑞𝑘,𝑙.(4.8) Realizing that 𝛿𝑖,𝑗𝑘,𝑙𝐈𝑑 represents the class of functions 𝐱 in 𝑥Ω whose values are zero on every dyadic point of 𝐷𝑁 except for 𝐱(𝑙2𝑘)=𝐈𝑑, {Δ𝑝,𝑞𝑘,𝑙}(𝑝,𝑞)𝑁 clearly appear as the coefficients of the decomposition of such functions in the bases 𝝍𝑝,𝑞 for (𝑝,𝑞) in 𝑁.

Denoting Ξ={Ξ𝑛,𝑘}(𝑛,𝑘)𝐼, a set of independent Gaussian variables of law 𝒩(𝟎,𝐈𝑑) on (Ω,,𝐏), and for all 𝑁>0, we form the finite dimensional Gauss-Markov vector [𝐗𝑁𝑖,𝑗](𝑖,𝑗)𝑁 as𝐗𝑁𝑖,𝑗=(𝑛,𝑘)𝑁𝝍𝑛,𝑘𝑚𝑖,𝑗𝚵𝑛,𝑘,(4.9) which, from Corollary 2.5, has the same law as [𝐗𝑡]𝑡𝐷𝑁, the finite-dimensional random vector obtained from sampling 𝐗 on 𝐷𝑁 (modulo a permutation on the indices). We then prove the following lemma that sheds light on the meaning of the construction.

Lemma 4.1. The Cholesky decomposition of the finite-dimensional covariance block matrix 𝚺𝑁 is given by 𝚺𝑁=Ψ𝑁Ψ𝑁𝑇.

Proof. For every 0𝑡,𝑠1, we compute the covariance of the finite-dimensional process 𝐗𝑁 as 𝐂𝑁(𝑡,𝑠)=𝔼𝐗𝑁𝑡𝐗𝑁𝑠𝑇=𝑁𝑛=00𝑘<2𝑛1𝝍𝑛,𝑘(𝑡)𝝍𝑛,𝑘(𝑠)𝑇.(4.10) From there, we write the finite-dimensional covariance block matrix 𝚺𝑁 in the recursively ordered basis 𝐟𝑖,𝑗 for 0𝑖𝑁, 0𝑗<2𝑖1, as 𝚺𝑁𝑖,𝑗𝑘,𝑙=𝐂𝑁𝑚𝑖,𝑗,𝑚𝑘,𝑙=𝑁𝑛=00𝑘<2𝑛1𝚿𝑖,𝑗𝑛,𝑘𝝍𝑘,𝑙𝑛,𝑘.(4.11) We already established that the matrix Ψ𝑁 was triangular with positive diagonal coefficient, which entails that the preceding equality provides us with the Cholesky decomposition of 𝚺.

In the finite-dimensional case, the inverse covariance or potential matrix is a well-defined quantity and we straightforwardly have the following corollary.

Corollary 4.2. The Cholesky decomposition of the finite-dimensional inverse covariance matrix 𝚺1𝑁 is given by 𝚺1𝑁=Δ𝑁𝑇Δ𝑁.

Proof. The result stems for the equalities 𝚺1𝑁=(Ψ𝑁Ψ𝑁𝑇)1=(Ψ1𝑁)𝑇Ψ1𝑁=Δ𝑁𝑇Δ𝑁.

4.1.2. The Lévy-Ciesielski Expansion

We now show that, asymptotically, the bases 𝝍𝑛,𝑘 allow us to faithfully build the Gauss-Markov process from which we have derived its expression. In this perspective we consider Ξ={Ξ𝑛,𝑘}(𝑛,𝑘), a set of independent Gaussian variables of law 𝒩(𝟎,𝐈𝑑) on (Ω,,𝐏), and, for all 𝑁>0, we form the finite-dimensional continuous Gaussian process 𝐙𝑁, defined for 0𝑡1 by𝐗𝑁𝑡=(𝑛,𝑘)𝑁𝝍𝑛,𝑘(𝑡)𝚵𝑛,𝑘,(4.12) which, from the result of Theorem 2.4, has the same law 𝐙𝑁𝑡=𝔼[𝑋𝑡𝑁]. We prove the following lemma.

Lemma 4.3. The sequence of processes 𝐗𝑁 almost surely converges towards a continuous Gaussian process denoted by 𝐗.

Proof. For all fixed 𝑁>0 and for any 𝜔 in Ω, we know that 𝑡𝐗𝑁𝑡(𝜔) is continuous. Moreover, we have established, that, for every 𝝃 in 𝜉Ω, 𝐗𝑁(𝝃) converges uniformly in 𝑡 toward a continuous limit denoted by 𝐗𝑁(𝝃). Therefore, in order to prove that lim𝑁𝐗𝑁 defines almost surely a process 𝐗 with continuous paths, it is sufficient to show that 𝐏𝜉(𝜉Ω)=1, where 𝐏𝜉=𝐏Ξ1 is the Ξ-induced measure on 𝜉Ω, which stems from a classical Borel-Cantelli argument. For 𝜉 a random variable of normal law 𝒩(0,1), and 𝑎>0, we have 𝐏||𝜉||>𝑎=2𝜋𝑎𝑒𝑢2/2𝑑𝑢2𝜋𝑎𝑢𝑎𝑒𝑢2/2𝑑𝑢=2𝜋𝑒𝑎2/2𝑎.(4.13) Then, for any 𝛿>0𝐏𝜉max0𝑘<2𝑛1||𝝃𝑛,𝑘||>2𝑛𝛿/2𝑑2𝑛𝐏||𝜉||>𝑑2𝑛𝛿/2=2𝜋2(1𝛿/2)𝑛exp2𝑛𝛿1.(4.14) Since the series 𝑛=02𝜋2(1𝛿/2)𝑛exp2𝑛𝛿1(4.15) is convergent, the Borel-Cantelli argument implies that 𝐏𝜉(𝜉Ω)=1. Eventually, the continuous almost-sure limit process 𝐗𝑡 is Gaussian as a countable sum of Gaussian processes.

Now that these preliminary remarks have been made, we can evaluate, for any 𝑡 and 𝑠 in [0,1], the covariance of 𝐗 as the limit of the covariance of 𝐗𝑁.

Lemma 4.4. For any 0𝑡,𝑠1, the covariance of 𝐗={𝐗𝑡=Ψ𝑡Ξ;0𝑡1} is 𝐂(𝑡,𝑠)=𝔼𝐗𝑡𝐗𝑠𝑇=𝐠(𝑡)𝐡(𝑡𝑠)𝐠(𝑠)𝑇.(4.16)

Proof. As Ξ𝑛,𝑘 are independent Gaussian random variables of normal law 𝒩(𝟎,𝐈𝑑), we see that the covariance of 𝐗𝑁 is given by 𝐂𝑁(𝑡,𝑠)=𝔼𝐗𝑁𝑡𝐗𝑁𝑠𝑇=(𝑛,𝑘)𝑁𝝍𝑛,𝑘(𝑡)𝝍𝑛,𝑘(𝑠)𝑇.(4.17) To compute the limit of the right-hand side, we need to remember that the element of the bases 𝝍𝑛,𝑘 and the functions 𝜙𝑛,𝑘 are linked by the following relation: 𝝍𝑛,𝑘(𝑡)=𝒦𝜙𝑛,𝑘=𝐠(𝑡)𝑈𝟙[0,𝑡](𝑠)𝐟(𝑠)𝜙𝑛,𝑘(𝑠)𝑑𝑠,(4.18) from which we deduce 𝐂𝑁(𝑡,𝑠)=𝐠(𝑡)(𝑛,𝑘)𝑁𝑈𝟙[0,𝑡](𝑢)𝐟(𝑢)𝜙𝑛,𝑘(𝑢)𝑑𝑢𝑈𝟙[0,𝑠](𝑣)𝐟(𝑣)𝜙𝑛,𝑘(𝑣)𝑑𝑣𝑇𝐠(𝑠)𝑇.(4.19) Defining the auxiliary 𝑑×𝑑-valued function 𝜿𝑛,𝑘(𝑡)=𝑈𝟙[0,𝑡](𝑢)𝐟(𝑢)𝜙𝑛,𝑘(𝑢)𝑑𝑢,(4.20) we observe that the (𝑖,𝑗)-coefficient function reads 𝜿𝑛,𝑘𝑖,𝑗(𝑡)=𝑈𝟙[0,𝑡](𝑢)𝑙𝑖(𝐟(𝑢))𝑇𝑐𝑗𝜙𝑛,𝑘(𝑢)𝑑𝑢=𝑈𝟙[0,𝑡](𝑢)𝑐𝑖𝐟𝑇(𝑢)𝑇𝑐𝑗𝜙𝑛,𝑘(𝑢)𝑑𝑢,(4.21) where 𝟙[0,𝑡] is the real function that is one if 0𝑢𝑡 and zero otherwise. As we can write 𝟙[0,𝑡](𝑢)𝑐𝑖𝐟𝑇(𝑢)=𝐟𝑇(𝑢)0𝟙[0,𝑡](𝑢)0𝑖,(4.22) we see that the function 𝐟𝑖,𝑡=𝟙[0,𝑡]𝑐𝑖(𝐟𝑇) belongs to 𝐿2𝐟, so that we can write (𝜿𝑛,𝑘)𝑖,𝑗(𝑡) as a scalar product in the Hilbert space 𝐿2𝐟: 𝜿𝑛,𝑘𝑖,𝑗(𝑡)=𝑈𝐟𝑇𝑖,𝑡(𝑢)𝑐𝑗𝜙𝑛,𝑘(𝑢)𝑑𝑢=𝐟𝑖,𝑡,𝑐𝑗𝜙𝑛,𝑘.(4.23) We then specify the (𝑖,𝑗)-coefficient of 𝐠1(𝑡)𝐂𝑁(𝑡,𝑠)(𝐠1(𝑠))𝑇 writing (𝑛,𝑘)𝑁𝜿(𝑡)𝜿(𝑠)𝑇𝑖,𝑗=(𝑛,𝑘)𝑁𝑑1𝑝=0𝐟𝑖,𝑡,𝑐𝑗𝜙𝑛,𝑘𝐟𝑗,𝑠,𝑐𝑗𝜙𝑛,𝑘,(4.24) and, remembering that the family of functions 𝑐𝑗(𝜙𝑛,𝑘) forms a complete orthonormal system of 𝐿2𝐟, we can use the Parseval identity, which reads (𝑛,𝑘)𝜿(𝑡)𝜿(𝑠)𝑇𝑖,𝑗=𝐟𝑖,𝑡,𝐟𝑗,𝑠=𝑈𝟙[0,𝑡](𝑢)𝑐𝑖𝐟𝑇(𝑢)𝑇𝟙[0,𝑠](𝑢)𝑐𝑗𝐟𝑇(𝑢)𝑑𝑢=𝑡𝑠0𝐟𝐟𝑇𝑖,𝑗(𝑢)𝑑𝑢.(4.25) Thanks to this relation, we can conclude the evaluation of the covariance since lim𝑁𝐂𝑁(𝑡,𝑠)=𝐠(𝑡)𝑡𝑠0𝐟𝐟𝑇(𝑢)𝑑𝑢𝐠(𝑠)𝑇=𝐠(𝑡)𝐡(𝑡𝑠)𝐠(𝑠)𝑇.(4.26)

We stress the fact that the relation𝐂(𝑡,𝑠)=(𝑛,𝑘)𝝍𝑛,𝑘(𝑡)𝝍𝑛,𝑘(𝑠)𝑇=𝚿(𝑡)𝚿𝑇(𝑠)(4.27) provides us with a continuous version of the Cholesky decomposition of the covariance kernel 𝐂. Indeed, if we chose 𝝈𝑛,𝑘 as the Cholesky square root of 𝚺𝑛,𝑘, we remark that the operators Φ are triangular in the following sense: consider the chain of nested vectorial spaces {𝐹𝑛,𝑘}(𝑛,𝑘)𝐹0,0𝐹1,0𝐹2,0𝐹2,1𝐹𝑛,0𝐹𝑛,2𝑛1𝜉Ω(4.28) with 𝐹𝑛,𝑘=span{𝐟𝑖,𝑗0𝑖𝑛,0𝑗𝑘}; then, for every (𝑛,𝑘) in , the operator Ψ transforms the chain {𝐹𝑛,𝑘}(𝑛,𝑘) into the chain0,01,02,02,1𝑛,0𝑛,2𝑛1𝑥Ω(4.29) with 𝑛,𝑘=span{Ψ𝑖,𝑗0𝑖𝑛,0𝑗𝑘}.

The fact that this covariance is equal to the covariance of the process 𝐗, solution of (2.1) implies that we have the following fundamental result.

Theorem 4.5. The process 𝐗 is equal in law to the initial Gauss-Markov process 𝐗 used to construct the basis of functions.

Remark 4.6. Our multiresolution representation of the Gauss-Markov processes appears to be the direct consequence of the fact that, because of the Markov property, the Cholesky decomposition of the finite-dimensional covariance admits a simple inductive continuous limit. More generally, triangularization of the kernel operators has been studied in depth [28, 4042], and it would be interesting to investigate if these results make possible a similar multiresolution approach for non-Markov Gaussian processes. In this regard, we naturally expect to lose the compactness of the supports of a putative basis.

Remark 4.7. We eventually underline the fact that large deviations related to this convergence can be derived through the use of the Baldi and Caramellino good rate function related to the Gaussian pinned processes [43, 44].

4.2. Optimality Criterion of the Decomposition

In the following, we draw from the theory of interpolating splines to further characterize the nature of our proposed basis for the construction of the Gauss-Markov processes. Essentially adapting the results from the previous works [45, 46], we first show that the finite-dimensional sample paths of our construction induce a nested sequence 𝑁 of the reproducing Hilbert kernel space (RKHS). In turn, the finite-dimensional process 𝐗𝑁 naturally appears as the orthogonal projection of the infinite-dimensional process 𝐗 onto 𝑁. We then show that such an RKHS structure allows us to define a unicity criterion for the finite-dimensional sample path as the only functions of that minimize a functional, called Dirichlet energy, under constraint of interpolation on 𝐷𝑁 (equivalent to conditioning on the times 𝐷𝑁). In this respect, we point out that the close relation between the Markov processes and the Dirichlet forms is the subject of a vast literature, largely beyond the scope of the present paper (see, e.g., [18]).

4.2.1. Sample Paths Space as a Reproducing Hilbert Kernel Space

In order to define the finite-dimensional sample paths as a nested sequence of RKHSs, let us first define the infinite-dimensional operator𝚽𝑙2𝜉Ω𝐿2𝐟,𝝃𝚽[𝝃]=𝑡(𝑛,𝑘)𝜙𝑛,𝑘(𝑡)𝝃𝑛,𝑘.(4.30) Since we know that the column functions of 𝜙𝑛,𝑘 form a complete orthonormal system of 𝐿2𝐟, the operator Φ is an isometry and its inverse satisfies Φ1=Φ𝑇, which reads for all 𝐯 in 𝐿2𝐟𝚽1[𝐯]𝑛,𝑘=𝑈𝜙𝑇𝑛,𝑘(𝑡)𝐯(𝑡)𝑑𝑡=𝒫𝜙𝑛,𝑘,𝐯.(4.31) Equipped with this infinite-dimensional isometry, we then consider the linear operator =ΦΔ suitably defined on the set=𝐮𝐶0𝑈,𝑑[𝐮]𝐿2𝐟=𝐮𝐶0𝑈,𝑑𝚫[𝐮]𝑙2𝜉Ω(4.32) with 𝝃22=𝑛,𝑘|𝝃𝑛,𝑘|22, the 𝑙2 norm of 𝜉Ω. The set form an infinite-dimensional vectorial space that is naturally equipped with the inner product(𝐮,𝐯)2,𝐮,𝐯=𝑈[𝐮](𝑡)𝑇[𝐯](𝑡)𝑑𝑡=([𝐮],[𝐯]).(4.33) Moreover since 𝐮(0)=𝐯(0)=𝟎, such an inner product is definite positive and, consequently, forms an Hilbert space.

Remark 4.8. Two straightforward remarks are worth making. First, the space is strictly included in the infinite-dimensional sample paths space 𝑥Ω. Second, notice that, in the favorable case 𝑚=𝑑, if 𝐟 is everywhere invertible with continuously differentiable inverse, we have =𝒟=𝒦1. More relevantly, the operator can actually consider a first-order differential operator from to 𝐿2𝐟 as a general left inverse of the integral operator 𝒦. Indeed, realizing that on 𝐿2𝐟,𝒦 can be expressed as 𝒦=ΨΦ1, we clearly have 𝒦=𝚽𝚫𝚿𝚽1=𝐼𝑑𝐿2𝐟.(4.34)

We know motivate the introduction of the Hilbert space by the following claim.

Proposition 4.9. The Hilbert space (,,) is a reproducing kernel Hilbert space (RKHS) with 𝑑×𝑑-valued reproducing kernel 𝐂, the covariance function of the process 𝐗.

Proof. Consider the problem of finding all elements 𝐮 of solution of the equation [𝐮]=𝐯 for 𝐯 in 𝐿2𝐟. The operator 𝒦 provides us with a continuous 𝑑×𝑚-valued kernel function 𝐤: (𝑡,𝑠)𝑈2,𝐤(𝑡,𝑠)=𝟙[0,𝑡](𝑠)𝐠(𝑡)𝐟(𝑠),(4.35) which is clearly the Green function for our differential equation. This entails that the following equalitiy holds for every 𝐮 in : 𝐮(𝑡)=𝑈𝐤(𝑡,𝑠)[𝐮](𝑠)𝑑𝑠.(4.36) Moreover, we can decompose the kernel 𝐤 in the 𝐿2𝐟 sense as 𝐤(𝑡,𝑠)=(𝑛,𝑘)𝝍𝑛,𝑘(𝑡)𝜙𝑇𝑛,𝑘(𝑠)(4.37) since we have 𝐤(𝑡,𝑠)=𝒦𝛿𝑠𝐼𝑑𝐿2𝐟(𝑡)=𝒦(𝑛,𝑘)𝜙𝑛,𝑘𝜙𝑇𝑛,𝑘(𝑠)(𝑡)=(𝑛,𝑘)𝒦𝜙𝑛,𝑘(𝑡)𝜙𝑇𝑛,𝑘(𝑠),(4.38) with 𝛿𝑠=𝛿(𝑠). Then, we clearly have 𝐂(𝑡,𝑠)=𝑈𝐤(𝑡,𝑢)𝐤(𝑠,𝑢)𝑇𝑑𝑢=(𝑛,𝑘)𝝍𝑛,𝑘(𝑡)𝝍𝑇𝑛,𝑘(𝑠),(4.39) where we recognize the covariance function of 𝐗, which implies 𝐤(𝑡,𝑠)=(𝑛,𝑘)𝝍𝑛,𝑘(𝑡)𝜙𝑛,𝑘𝑇(𝑠)=[𝐂(𝑡,)].(4.40) Eventually, for all 𝐮 in 𝐿2𝐟, we have 𝐮(𝑡)=𝑈[𝐂(𝑡,)](𝑠)[𝐮](𝑠)𝑑𝑠=𝒫𝐂(𝑡,),𝐮,(4.41) where we have introduced the 𝒫-operator associated with the inner product ,: for all 𝑑×𝑑-valued functions 𝐀 and 𝐁 defined on 𝑈 such that the columns 𝑐𝑖(𝐀) and 𝑐𝑖(𝐁), 0𝑖<𝑑, are in , we define the matrix 𝒫𝐀,𝐁 in 𝑑×𝑑 by 0𝑖,𝑗<𝑑,𝒫𝐀,𝐁𝑖,𝑗=𝑐𝑖(𝐀),𝑐𝑗(𝐁).(4.42) By the Moore-Aronszajn theorem [47], we deduce that there is a unique reproducing kernel Hilbert space associated with a given covariance kernel. Thus, is the reproducing subspace of 𝐶0(𝑈,𝑑) corresponding to the kernel 𝐂, with respect to the inner product ,.

Remark 4.10. From a more abstract point of view, it is well known that the covariance operator of a Gaussian measure defines an associated Hilbert structure [48, 49].

In the sequel, we will use the space as the ambient Hilbert space to define the finite-dimensional sample paths spaces as a nested sequence of RKHS. More precisely, let us write for 𝑁 the finite-dimensional subspace of 𝑁=𝐮𝐶0𝑈,𝑑[𝐮]𝐿2𝐟,𝑁,(4.43) with the space 𝐿2𝐟,𝑁 being defined as𝐿2𝐟,𝑁=span𝑐𝑖(𝜙𝑛,𝑘)𝑛,𝑘𝑁,0𝑖<𝑑.(4.44)

We refer to such spaces as finite-dimensional approximation spaces since we remark that𝑁=span𝑐𝑖(𝝍𝑛,𝑘)𝑛,𝑘𝑁,0𝑖<𝑑=𝚿𝑁𝜉Ω𝑁,(4.45) which means that the space 𝑁 is made of the sample space of the finite-dimensional process 𝐗𝑁. The previous definition makes obvious the nested structure 01, and it is easy to characterize each space 𝑁 as a reproducing Hilbert kernel space.

Proposition 4.11. The Hilbert spaces (𝑁,,) are reproducing kernel Hilbert spaces (RKHSs) with 𝑑×𝑑-valued reproducing kernel 𝐂𝑁, the covariance function of the process 𝐗𝑁.

Proof. The proof this proposition follows the exact same argument as that in the case of , but with the introduction of finite-dimensional kernels 𝐤𝑁(𝑡,𝑠)𝑈2,𝐤𝑁(𝑡,𝑠)=(𝑛,𝑘)𝑁𝝍𝑛,𝑘(𝑡)𝜙𝑇𝑛,𝑘(𝑠),(4.46) and the corresponding covariance function (𝑡,𝑠)[0,1]2,𝐂𝑁(𝑡,𝑠)=(𝑛,𝑘)𝑁𝝍𝑛,𝑘(𝑡)𝝍𝑇𝑛,𝑘(𝑠).(4.47)

4.2.2. Finite-Dimensional Processes as Orthogonal Projections

The framework set in the previous section offers a new interpretation of our construction. Indeed, for all 𝑁>0, the columns of {𝝍𝑛,𝑘}(𝑛,𝑘)𝑁 form an orthonormal basis of the space 𝑁:𝒫𝝍𝑛,𝑘,𝝍𝑝,𝑞=𝒫𝝍𝑛,𝑘,𝝍𝑝,𝑞=𝒫𝜙𝑛,𝑘,𝜙𝑝,𝑞=𝛿𝑛,𝑘𝑝,𝑞.(4.48) This leads to defining the finite-dimensional approximation 𝐱𝑁 of an sample path 𝐱 of as the orthogonal projection of 𝐱 on 𝑁 with respect to the inner product ,. At this point, it is worth remembering that the space is strictly contained in 𝑥Ω and does not coincide with 𝑥Ω: actually one can easily show that 𝐏()=𝟎. We devote the rest of this section to defining the finite-dimensional processes 𝐙𝑁=𝔼𝑁[𝑋] resulting from the conditioning on 𝐷𝑁, as pathwise orthogonal projection of the original process 𝐗 on the sample space 𝑁.

Proposition 4.12. For any 𝑁>0, the conditioned processes 𝔼𝑁[𝐗] can be written as the orthogonal projection of 𝐗 on 𝑁 with respect to ,: 𝔼𝑁[𝐗]=(𝑛,𝑘)𝑁𝝍𝑛,𝑘𝒫𝝍𝑛,𝑘,𝐗.(4.49)

The only hurdle to prove Proposition 4.12 is purely technical in the sense that the process 𝐗 exists in a larger space than : we need to find a way to extend the definition of , so that the expression bears a meaning. Before answering this point quite straightforwardly, we need to establish the following lemma.

Lemma 4.13. Writing the Gauss-Markov process 𝐗𝑡=10𝐤(𝑡,𝑠)𝑑𝐖𝑠, for all 𝑁>0, the conditioned process 𝐙𝑁=𝔼𝑁[𝐗] is expressed as the stochastic integral 𝐙𝑁=10𝐤𝑁(𝑡,𝑠)𝑑𝐖𝑠with𝐤𝑁(𝑡,𝑠)=(𝑛,𝑘)𝑁𝝍𝑛,𝑘(𝑡)𝜙𝑇𝑛,𝑘(𝑠).(4.50)

Proof. In the previous section, we have noticed that the kernel 𝐤𝑁 converges toward the kernel 𝐤 (the Green function) in the 𝐿2𝐟 sense 𝐤(𝑡,𝑠)=(𝑛,𝑘)𝝍𝑛,𝑘(𝑡)𝜙𝑇𝑛,𝑘(𝑠)=lim𝑁(𝑛,𝑘)𝑁𝝍𝑛,𝑘(𝑡)𝜙𝑇𝑛,𝑘(𝑠)𝐤𝑁=lim𝑁𝐤𝑁(𝑡,𝑠).(4.51) This implies that the process 𝐗 as the stochastic integral can also be written as 𝐗𝑡=𝐠(𝑡)𝑈𝟙[0,𝑡](𝑠)(𝑠)𝑑𝐖𝑠=10𝐤(𝑡,𝑠)𝑑𝐖𝑠=lim𝑁10𝐤𝑁(𝑡,s)𝑑𝐖𝑠.(4.52) Specifying the decomposition of 𝐤𝑁, we can then naturally express 𝐗 as the convergent sum 𝐗𝑡=(𝑛,𝑘)𝝍𝑛,𝑘𝚵𝑛,𝑘with𝚵𝑛,𝑘=10𝜙𝑇𝑛,𝑘(𝑠)𝑑𝐖𝑠,(4.53) where the orthonormality property of the 𝜙𝑛,𝑘 with respect to (,) makes the vectors Ξ𝑛,𝑘 appear as independent 𝑑-dimensional Gaussian variables of law 𝒩(0,𝐈𝑑). It is then easy to see that by definition of the elements 𝝍𝑛,𝑘, for almost every 𝜔 in Ω, we then have 𝑁>0,0𝑡1,𝐙𝑁(𝜔)=𝔼𝑁[𝐗](𝜔)=(𝑛,𝑘)𝑁𝝍𝑛,𝑘𝚵𝑛,𝑘(𝜔),(4.54) and we finally recognize in the previous expression that for all 0𝑡1𝐙𝑁𝑡=(𝑛,𝑘)𝑁𝝍𝑛,𝑘𝚵𝑛,𝑘=(𝑛,𝑘)𝑁𝝍𝑛,𝑘(𝑡)𝑈𝜙𝑇𝑛,𝑘(𝑠)𝑑𝐖s=10𝐤𝑁(𝑡,𝑠)𝑑𝐖𝑠.(4.55)

We can now proceed to justify the main result of Proposition 4.12.

Proof. The finite-dimensional processes 𝐙𝑁 defined through Lemma 4.13 have sample paths 𝑡𝐙𝑁𝑡(𝜔) belonging to 𝑁. Moreover, for almost every 𝜔 in Ω and for all 𝑛,𝑘 in 𝑁, 𝒫𝝍𝑛,𝑘,𝐙𝑁(𝜔)=𝒫𝝍𝑛,𝑘,10𝐤𝑁(𝑡,𝑠)𝑑𝐖𝑠(𝜔)=𝒫𝝍𝑛,𝑘,(𝑝,𝑞)𝐼𝑁𝝍𝑝,𝑞(𝜔)10𝜙𝑇𝑝,𝑞(𝑠)𝑑𝐖𝑠(𝜔)=10𝜙𝑇𝑛,𝑘(𝑠)𝑑𝐖𝑠(𝜔),(4.56) because of the orthonormality property of 𝝍𝑛,𝑘 with respect to ,. As the previous equalities hold for every 𝑁>0, the applications 𝐱𝒫𝝍𝑛,𝑘,𝐱 can naturally be extended on 𝑥Ω by continuity. Therefore, it makes sense to write, for all (𝑛,𝑘) in 𝑁, 𝒫𝝍𝑛,𝑘,𝐙𝑁=lim𝑁𝒫𝝍𝑛,𝑘,𝐙𝑁def=𝒫𝝍𝑛,𝑘,𝐗 even if 𝐗 is defined into a larger sample space than . In other words, we have 𝒫𝝍𝑛,𝑘,𝐗=10𝜙𝑇𝑛,𝑘(𝑠)𝑑𝐖𝑠=𝚵𝑛,𝑘,(4.57) and we can thus express the conditioned process 𝐙𝑁=𝔼𝑁[𝐗] as the orthogonal projection of 𝐗 onto the finite sample path 𝑁 by writing 𝐙𝑁=(𝑛,𝑘)𝑁𝝍𝑛,𝑘𝒫𝝍𝑛,𝑘,𝐗.(4.58)

4.2.3. Optimality Criterion of the Sample Paths

Proposition 4.12 elucidates the structure of the conditioned processes 𝐙𝑁 as pathwise orthogonal projections of 𝐗 on the finite-dimensional RKHS 𝑁. It allows us to cast the finite sample paths in a geometric setting and incidentally, to give a characterization of them as the minimizer of some functionals. In doing so, we shed a new light on well-known results of the interpolation theory [5052] and extend them to the multidimensional case.

The central point of this section reads as follows.

Proposition 4.14. Given a function 𝐱 in , the function 𝐱𝑁=(ΨΔ𝑁)[𝐱] belongs to 𝑁 and is defined by the following optimal criterion: 𝐱𝑁 is the only function in interpolating 𝐱 on 𝐷𝑁 such that the functional 𝐲,𝐲=𝐲(𝑡)22=10||𝐲(𝑡)||22𝑑𝑡(4.59) takes its unique minimal value over in 𝐱𝑁.

Proof. The space 𝑁 has been defined as 𝑁=Ψ𝑁[𝜉Ω𝑁]=ΨΔ𝑁[] so that, for all 𝐱 in , 𝐱𝑁 clearly belongs to 𝑁. Moreover, 𝐱𝑁 interpolates 𝐱 on 𝐷𝑁: indeed, we know that the finite-dimensional operators Δ𝑁 and Ψ1𝑁 are inverse of each other Δ𝑁=Ψ1𝑁, which entails that for all 𝑡 in 𝐷𝑁𝐱𝑁(𝑡)=𝚿𝚫𝑁[𝐱](𝑡)=𝚿𝑁𝚫𝑁[𝐱](𝑡)=𝐱(𝑡),(4.60) where we use the fact that, for any 𝝃 in 𝜉Ω and for all 𝑡 in 𝐷𝑁, Ψ𝑁[𝝃](𝑡)=Ψ[𝝃](𝑡) (recall that 𝝍𝑛,𝑘(𝑡)=𝟎 if 𝑛>𝑁 and 𝑡 belongs to 𝐷𝑁).
Let us now show that 𝐱𝑁 is determined in by the announced optimal criterion. Suppose 𝐲 belongs to and interpolates 𝐱 on 𝐷𝑁, and remark that we can write 𝐲,𝐲=[𝐲]22=(𝚽𝚫)[𝐲](𝑡)22=𝚫[𝐲]22,(4.61) since Φ is an isometry. Then, consider Δ[𝐲] in 𝑙2(𝜉Ω) and remark that, since for all (𝑛,𝑘) in 𝑁, 𝜹𝑛,𝑘 are Dirac measures supported by 𝐷𝑁, we have (𝑛,𝑘)𝑁,𝚫𝑛,𝑘𝐲=𝒫𝜹𝑛,𝑘,𝐲=𝒫𝜹𝑛,𝑘,𝐱=𝚫𝑛,𝑘[𝐱]=𝚫𝑛,𝑘𝐱𝑁.(4.62) This entails 𝚫𝐲22𝑑𝑡=(𝑛,𝑘)||𝚫𝑛,𝑘𝐲||22(𝑛,𝑘)𝑁||𝚫𝑛,𝑘𝐲||22=𝚫[𝐱𝑁]22𝑑𝑡.(4.63) Since, by definition of 𝐱𝑁, 𝜹𝑛,𝑘[𝐱𝑁]=𝟎 if 𝑛>𝑁. Moreover, the minimum 𝐱𝑁,𝐱𝑁 is only attained for 𝐲 such that 𝜹𝑛,𝑘[𝐲]=𝟎 if 𝑛>𝑁 and 𝜹𝑛,𝑘[𝐲]=𝜹𝑛,𝑘[𝐱] if 𝑛𝑁, which defines univocally 𝐱𝑁. This establishes that, for all 𝐲 in such that for all 𝑡 in 𝐷𝑁, 𝐲(𝑡)=𝐱(𝑡), we have 𝐱𝑁,𝐱𝑁𝐲,𝐲 and the equality case holds if and only if 𝐲=𝐱𝑁.

Remark 4.15. When represents a regular differential operator of order 𝑑, 𝑑𝑖=1𝑎𝑖(𝑡)𝐷𝑖, where 𝐷=𝑑/𝑑𝑡, that is, for 𝑑𝐗𝑡=𝜶(𝑡)𝐗𝑡+𝚪(𝑡)𝑑𝑊𝑡,(4.64) with 𝜶(𝑡)=011𝑎𝑑𝑎𝑑1𝑎1,𝚪(𝐭)=001.(4.65) The finite-dimensional sample paths coincide exactly with the spline interpolation of order 2𝑑+1, which are well known to satisfy the previous criterion [46]. This example will be further explored in the example section.

The Dirichlet energy simply appears as the squared norm induced on by the inner product ,, which in turn can be characterized as a Dirichlet quadratic form on . Actually, such a Dirichlet form can be used to define the Gauss-Markov process, extending the Gauss-Markov property to processes indexed on the multidimensional spaces parameter [19]. In particular, for an 𝑛-dimensional parameter space, we can condition such Gauss-Markov processes on a smooth 𝑛1-dimensional boundary. Within the boundary, the sample paths of the resulting conditioned process (the solution to the prediction problem in [19]) are the solutions to the corresponding Dirichlet problems for the elliptic operator associated with the Dirichlet form.

The characterization of the basis as the minimizer of such a Dirichlet energy (4.59) gives rise to an alternative method to compute the basis as the solution of a Dirichlet boundary value problem for an elliptic differential operator.

Proposition 4.16. Let us assume that 𝜶 and Γ are continuously differentiable and that Γ is invertible. Then, the functions 𝝁𝑛,𝑘 are defined as 𝝁𝑛,𝑘(𝑡)=𝝁𝑙(𝑡),𝑡𝑙𝑛,𝑘,𝑚𝑛,𝑘,𝝁𝑟(𝑡),𝑡𝑚𝑛,𝑘,𝑟𝑛,𝑘,𝟎,else,(4.66) where 𝝁𝑙 and 𝝁𝑟 are the unique solutions of the second-order 𝑑-dimensional linear differential equation 𝐮+𝚪1𝜶𝑇𝚪𝚪𝜶𝐮𝚪1𝜶𝑇𝚪𝚪𝜶+𝜶𝐮=𝟎(4.67) with the following boundary value conditions: 𝝁𝑙𝑙𝑛,𝑘=𝟎,𝝁𝑙𝑚𝑛,𝑘=𝐈𝑑,𝝁𝑟𝑚𝑛,𝑘=𝐈𝑑,𝝁𝑟𝑟𝑛,𝑘=𝟎.(4.68)

Proof. By Proposition 4.14, we know that 𝝁𝑛,𝑘(𝑡) minimizes the convex functional 10||[𝐮](𝑠)||22𝑑𝑠(4.69) over , being equal to zero outside the interval [𝑙𝑛,𝑘,𝑟𝑛,𝑘] and equal to one at the point 𝑡=𝑚𝑛,𝑘. Because of the hypotheses on 𝜶 and Γ, we have =𝒟 and we can additionally restrain our search to functions that are twice continuously differentiable. Incidentally, we only need to minimize separately the contributions on the interval [𝑙𝑛,𝑘,𝑚𝑛,𝑘] and [𝑚𝑛,𝑘,𝑟𝑛,𝑘]. On both intervals, this problem is a classical Euler-Lagrange problem (see, e.g., [53]) and is solved using basic principles of calculus of variations. We easily identify the Lagrangian of our problem as 𝐿𝑡,𝐮,𝐮=|||𝐮𝜶(𝑡)𝐮(𝑡)𝚪(𝐭)1|||22=𝐮(𝑡)𝜶(𝑡)𝐮(𝑡)𝑇(𝚪(𝑡))1𝐮(𝑡)𝜶(𝑡)𝐮(𝑡).(4.70) From there, after some simple matrix calculations, the Euler-Lagrange equations 𝜕𝐿(𝑡,𝐮,𝐮)𝜕𝑢𝑖𝑑𝑑𝑡𝜕𝐿𝑡,𝐮,𝐮𝜕𝑢𝑖=0,𝑖=1,,𝑑,(4.71) can be expressed under the form: 𝐮+𝚪1𝜶𝑇𝚪𝚪𝜶𝐮𝚪1𝜶𝑇𝚪Γ𝜶+𝜶𝐮=𝟎,(4.72) which ends the proof.

Remark 4.17. It is a simple matter of calculus to check that the expression of 𝝁 given in Proposition 2.1 satisfies (4.67). Notice also that, in the case Γ=𝐈𝑑, the differential equation becomes 𝐮+𝜶𝑇𝜶𝐮𝜶𝑇𝜶+𝜶𝐮=𝟎,(4.73) which is further simplified for constant 𝜶.

Under the hypotheses of Proposition 4.16, we can thus define 𝝁𝑛,𝑘 as the unique solution to the second-order linear differential equation (4.67) with the appropriate boundary values conditions. From this definition, it is then easy to derive the bases 𝝍𝑛,𝑘 by completing the following program.(1)Compute the 𝑡𝝁𝑛,𝑘(𝑡) by solving the linear ordinary differential problem.(2)Apply the differential operator 𝒟 to get the functions 𝒟[𝝁𝑛,𝑘].(3)Orthonormalize the column functions 𝑡𝑐𝑗(𝒟[𝝁𝑛,𝑘(𝑡)]) by the Gram-Schmidt process.(4)Apply the integral operator 𝒦 to get the desired functions 𝝍𝑛,𝑘 (or equivalently multiply the original function 𝑡𝝁𝑛,𝑘(𝑡) by the corresponding Gram-Schmidt triangular matrix).

Notice finally that each of these points is easily implemented numerically.

5. Examples: Derivation of the Bases for Some Classical Processes

5.1. One-Dimensional Case

In the one-dimensional case, the construction of the Gauss-Markov process is considerably simplified since we do not have to consider the potential degeneracy of matrix-valued functions. Indeed, in this situation, the centered Gauss-Markov process 𝑋 is solution of the one-dimensional stochastic equation𝑑𝑋𝑡=𝛼(𝑡)𝑋𝑡𝑑𝑡+Γ(𝑡)𝑑𝑊𝑡,(5.1) with 𝛼 homogeneously Hölder continuous and Γ positive continuous function. We then have the Doob representation𝑋𝑡=𝑔(𝑡)𝑡0𝑓(𝑠)𝑑𝑊𝑠,with𝑔(𝑡)=𝑒𝑡0𝛼(𝑣)𝑑𝑣,𝑓(𝑡)=Γ(𝑡)𝑒𝑡0𝛼(𝑣)𝑑𝑣.(5.2) Writing the function as(𝑡)=𝑡0𝑓2(𝑠)𝑑𝑠,(5.3) the covariance of the process reads for any 0𝑡,𝑠0𝐶(𝑡,𝑠)=𝑔(𝑡)𝑔(𝑠)(𝑡𝑠).(5.4) The variance of the Gauss-Markov bridge 𝐵𝑡 pinned in 𝑡𝑥 and 𝑡𝑧 yields𝜎𝑡𝑥,𝑡𝑧(𝑡)2=𝑔(𝑡)2(𝑡)𝑡𝑥𝑡𝑧(𝑡)𝑡𝑧𝑡𝑥.(5.5) These simple relations entail that the functions 𝜓𝑛,𝑘 are defined on their supports 𝑆𝑛,𝑘 by 𝜓𝑛,𝑘(𝑡)2=𝔼[(𝛿𝑛(𝑡))2] with𝔼(𝛿𝑛(𝑡))2=𝜎𝑙𝑛,𝑘,𝑟𝑛,𝑘(𝑡)2𝟙𝑆𝑛+1,2𝑘(𝑡)𝜎𝑙𝑛,𝑘,𝑚𝑛,𝑘(𝑡)2+𝟙𝑆𝑛+1,2𝑘+1(𝑡)𝜎𝑙𝑛,𝑘,𝑚𝑛,𝑘(𝑡)2.(5.6)

This reads on 𝑆𝑛+1,2𝑘𝜓𝑛,𝑘(𝑡)2=𝑔(𝑡)2(𝑡)𝑙𝑛,𝑘𝑟𝑛,𝑘(𝑡)𝑟𝑛,𝑘𝑙𝑛,𝑘(𝑡)𝑙𝑛,𝑘𝑚𝑛,𝑘(𝑡)𝑚𝑛,𝑘𝑙𝑛,𝑘(5.7) and on 𝑆𝑛+1,2𝑘+1 as𝜓𝑛,𝑘(𝑡)2=𝑔(𝑡)2(𝑡)𝑙𝑛,𝑘𝑟𝑛,𝑘(𝑡)𝑟𝑛,𝑘𝑙𝑛,𝑘(𝑡)𝑚𝑛,𝑘𝑟𝑛,𝑘(𝑡)𝑟𝑛,𝑘𝑚𝑛,𝑘(5.8) and therefore we have𝜓𝑛,𝑘(𝑡)=𝜎𝑛,𝑘𝑔(𝑡)(𝑡)𝑙𝑛,𝑘𝑔𝑚𝑛,𝑘𝑚𝑛,𝑘𝑙𝑛,𝑘,𝑙𝑛,𝑘𝑡𝑚𝑛,𝑘,𝜎𝑛,𝑘𝑔(𝑡)𝑟𝑛,𝑘(𝑡)𝑔𝑚𝑛,𝑘𝑟𝑛,𝑘𝑚𝑛,𝑘,𝑚𝑛,𝑘𝑡𝑟𝑛,𝑘,(5.9) with𝜎𝑛,𝑘=𝑟𝑛,𝑘𝑚𝑛,𝑘𝑚𝑛,𝑘𝑙𝑛,𝑘𝑟𝑛,𝑘𝑙𝑛,𝑘.(5.10) As for the first element, it simply results from the conditional expectation of the one-dimensional bridge pinned in 𝑙0,0=0 and 𝑟0,0=1:𝜓0,0(𝑡)=𝑔(𝑡)(𝑡)𝑙0,0𝑟0,0𝑙0,0.(5.11) In this class of processes, two paradigmatic process are the Wiener process and the Ornstein-Uhlenbeck processes with constant coefficients. In the case of the Wiener process, (𝑡)=𝑡 and 𝑔(𝑡)=1, which yields the classical triangular-shaped Schauder functions used by Lévy [8]. As for the Ornstein-Uhlenbeck process with constant coefficients 𝛼 and Γ, we have 𝑔(𝑡)=exp(𝛼𝑡), 𝑓(𝑡)=Γexp(𝛼𝑡) and (𝑡)=(Γ/2𝛼)(1𝑒2𝛼𝑡), yielding for the construction basis the expressions𝜓𝑛,𝑘(𝑡)=Γ𝛼sinh𝛼𝑡𝑙𝑛,𝑘sinh𝛼𝑚𝑛,𝑘𝑙𝑛,𝑘,𝑙𝑛,𝑘𝑡𝑚𝑛,𝑘,Γ𝛼sinh𝛼𝑟𝑛,𝑘𝑡sinh𝛼𝑚𝑛,𝑘𝑙𝑛,𝑘,𝑚𝑛,𝑘𝑡𝑟𝑛,𝑘,𝜓0,0(𝑡)=Γ𝛼𝑒𝛼/2sinh(𝛼𝑡)sinh(𝛼),(5.12) which were already evidenced in [54].

5.2. Multidimensional Case

In the multidimensional case, the explicit expressions for the basis functions 𝝍𝑛,𝑘 make fundamental use of the flow 𝐅 of the underlying linear equation (2.3) for a given function 𝜶. For commutative forms of 𝜶 (i.e., such that 𝜶(𝑡)𝜶(𝑠)=𝜶(𝑠)𝜶(𝑡) for all 𝑡,𝑠), the flow can be formally expressed as an exponential operator. It is, however, a notoriously difficult problem to find a tractable expression for general 𝛼. As a consequence, it is only possible to provide closed-from formulae for our basis functions in very specific cases.

5.2.1. Multidimensional Gauss-Markov Rotations

We consider in this section that 𝜶 is antisymmetric and constant and Γ𝑑×𝑚 such that Γ=𝜎2𝐈𝑑. For 𝜶 antisymmetric, since 𝜶𝑇(𝑡)=𝜶(𝑡), we have𝐅(𝑠,𝑡)𝑇=𝐅(𝑠,𝑡)1,(5.13) that is, the flow is unitary. This property implies that𝐡𝑢(𝑠,𝑡)=𝜎2𝑡𝑠𝐅(𝑤,𝑢)𝐅(𝑤,𝑢)𝑇𝑑𝑤=𝜎2(𝑡𝑠)𝐈𝑑,(5.14) which yields by definition of 𝝈𝑛,𝑘𝝈𝑛,𝑘𝝈𝑇𝑛,𝑘=𝜎2𝑚𝑛,𝑘𝑙𝑛,𝑘𝑟𝑛,𝑘𝑚𝑛,𝑘𝑟𝑛,𝑘𝑙𝑛,𝑘𝐈𝑑.(5.15) The square root 𝝈𝑛,𝑘 is then uniquely defined (by choosing both Cholesky and symmetrical square roots) by𝝈𝑛,𝑘=𝜎𝑚𝑛,𝑘𝑙𝑛,𝑘𝑟𝑛,𝑘𝑚𝑛,𝑘𝑟𝑛,𝑘𝑙𝑛,𝑘𝐈𝑑,(5.16) and 𝝍𝑛,𝑘(𝑡) reads𝝍𝑛,𝑘(𝑡)=𝜎𝑟𝑛,𝑘𝑚𝑛,𝑘𝑚𝑛,𝑘𝑙𝑛,𝑘𝑟𝑛,𝑘𝑙𝑛,𝑘𝑡𝑙𝑛,𝑘𝐅𝑚𝑛,𝑘,𝑡,𝑙𝑛,𝑘𝑡𝑚𝑛,𝑘,𝜎𝑚𝑛,𝑘𝑙𝑛,𝑘𝑟𝑛,𝑘𝑚𝑛,𝑘𝑟𝑛,𝑘𝑙𝑛,𝑘𝑟𝑛,𝑘𝑡𝐅𝑚𝑛,𝑘,𝑡,𝑙𝑛,𝑘𝑡𝑚𝑛,𝑘.(5.17)

Recognizing the (𝑛,𝑘) element of the Schauder basis for the construction of the one-dimensional Wiener process𝑠𝑛,𝑘(𝑡)=𝑟𝑛,𝑘𝑚𝑛,𝑘𝑟𝑛,𝑘𝑙𝑛,𝑘𝑚𝑛,𝑘𝑙𝑛,𝑘𝑡𝑙𝑛,𝑘,𝑙𝑛,𝑘𝑡𝑚𝑛,𝑘,𝑚𝑛,𝑘𝑙𝑛,𝑘𝑟𝑛,𝑘𝑙n,𝑘𝑟𝑛,𝑘𝑚𝑛,𝑘𝑟𝑛,𝑘𝑡,𝑙𝑛,𝑘𝑡𝑚𝑛,𝑘,(5.18) we obtain the following formula:𝝍𝑛,𝑘(𝑡)=𝜎𝑠𝑛,𝑘(𝑡)𝐅𝑡𝑚𝑛,𝑘.(5.19)

This form shows that the Schauder basis for multidimensional rotations results from the multiplication of the triangular-shaped elementary function used for the Lévy-Ciesielski construction of the Wiener process with the flow of the equation, that is, the elementary rotation.

The simplest example of this kind is the stochastic sine and cosine process corresponding to𝜶=0110,𝚪=𝜎2𝐈2.(5.20) In that case, 𝝍𝑛,𝑘 has the expression𝝍𝑛,𝑘(𝑡)=𝑠𝑛,𝑘(𝑡)cos𝑡𝑚𝑛,𝑘sin𝑡𝑚𝑛,𝑘sin𝑡𝑚𝑛,𝑘cos𝑡𝑚𝑛,𝑘.(5.21) Interestingly, the different basis functions have the structure of the solutions of the nonstochastic oscillator equation. One of the equations perturbs the trajectory in the radial component of the deterministic solution and the other one in the tangential direction. We represent such a construction in Figure 2 with the additional conditioning that 𝐗1=𝐗0, that is, imposing that the trajectory forms a loop between time 0 and 1.

5.2.2. The Successive Primitives of the Wiener Process

In applications, it often occurs that people use smooth stochastic processes to model the integration of noisy signals. This is for instance the case of a particular subject of a Brownian forcing or of the synaptic integration of noisy inputs [55]. Such smooth processes involves in general integrated martingales, and the simplest example of such processes are the successive primitives of a standard Wiener process.

Let 𝑑>2, and denote by 𝑋𝑑𝑡 the 𝑑1th order primitive of the Wiener process. This process can be defined via the lower-order primitives 𝑋𝑘𝑡 for 𝑘<𝑑 via the relations𝑑𝑋𝑘+1𝑡=𝑋𝑘𝑡𝑑𝑡,𝑘<𝑑,𝑑𝑋1𝑡=𝑑𝑊𝑡,(5.22) where 𝑊𝑡 is a standard real Wiener process. These equations can be written in our formalism as𝑑𝐗𝑡=𝜶(𝑡)𝐗𝑡+𝚪(𝑡)𝑑𝑊𝑡,(5.23) with𝜶(𝑡)=0110,𝚪(𝐭)=001.(5.24) In particular, though none of the integrated processes 𝑋𝑘 for 𝐾>1 is Markov by itself, the 𝑑-uplet 𝐗=(𝑋𝑑,,𝑋1) is a Gauss-Markov process.

Furthermore, because of the simplicity and the sparsity of the matrices involved, we can identify in a compact form all the variables used in the computation of the construction basis for these processes. In particular, the flow 𝐅 of the equation is the exponential of the matrix 𝜶, and since 𝛼 is nilpotent, it is easy to show that 𝐅 has the expression,𝐅(𝑠,𝑡)=1(𝑡𝑠)(𝑡𝑠)22(𝑡𝑠)𝑑1(𝑑1)!(𝑡𝑠)22(𝑡𝑠)1(5.25) and the only nonzero entry of the 𝑑×𝑑 matrix Γ is one at position (𝑑1,𝑑1). Using this expression and the highly simple expression of Γ, we can compute the general element of the matrix 𝐡𝑢(𝑡,𝑠), which reads𝐡𝑢(𝑠,𝑡)𝑖,𝑗=(1)𝑖+𝑗(𝑡𝑢)2𝑑1(𝑖+𝑗)(𝑠𝑢)2𝑑1(𝑖+𝑗)(2𝑑1(𝑖+𝑗))(𝑑1𝑖)!(𝑑1𝑗)!.(5.26) Eventually, we observe that the functions 𝝍𝑛,𝑘, yielding the multiresolution description of the integrated Wiener processes, are directly deduced from the matrix-valued function𝐜𝑛,𝑘(𝑡)𝑖,𝑗=𝝍𝑛,𝑘𝐋1𝑛,𝑘=𝐠(𝑡)𝐡𝑙𝑛,𝑘,𝑡,𝑙𝑛,𝑘𝑡𝑚𝑛,𝑘,𝝍𝑛,𝑘𝐑1𝑛,𝑘=𝐠(𝑡)𝐡𝑡,𝑟𝑛,𝑘,𝑚𝑛,𝑘𝑡𝑟𝑛,𝑘,(5.27) whose components are further expressed as𝐜𝑛,𝑘(𝑡)𝑖,𝑗=𝑑1𝑝=𝑖(1)𝑝+𝑗𝑡𝑖𝑝(𝑖𝑝)!𝑡2𝑑1(𝑝+𝑗)𝑙2𝑑1(𝑝+𝑗)𝑛,𝑘(2𝑑1(𝑝+𝑗))(𝑑1𝑝)!(𝑑1𝑗)!,(5.28) for 𝑙𝑛,𝑘𝑡𝑚𝑛,𝑘 and as𝐜𝑛,𝑘(𝑡)𝑖,𝑗=𝑑1𝑝=𝑖(1)𝑝+𝑗𝑡𝑖𝑝(𝑖𝑝)!𝑚2𝑑1(𝑝+𝑗)𝑛,𝑘𝑡2𝑑1(𝑝+𝑗)(2𝑑1(𝑝+𝑗))(𝑑1𝑝)!(𝑑1𝑗)!,(5.29) for 𝑚𝑛,𝑘𝑡𝑟𝑛,𝑘. The final computation of the 𝝍𝑛,𝑘 involves the computation of 𝐋𝑛,𝑘 and 𝐑𝑛,𝑘, which in the general case can become very complex. However, this expression is highly simplified when assuming that 𝑚𝑛,𝑘 is the middle of the interval [𝑙𝑛,𝑘,𝑟𝑛,𝑘]. Indeed, in that case, we observe that, for any (𝑖,𝑗) such that 𝑖+𝑗 is odd, (𝐡𝑚(𝑙,𝑟))𝑖,𝑗=𝟎, which induces the same property on the covariance matrix 𝚺𝑛,𝑘 and on the polynomials (𝐜𝑛,𝑘(𝑡))𝑖,𝑗. This property gives therefore a preference to the dyadic partition that provides simple expressions for the basis elements in any dimensions, and allows simple computations of the basis.

Remark 5.1. Observe that, for all 0𝑖<𝑑1, we have 𝐜𝑛,𝑘(𝑡)𝑖,𝑗=𝐜𝑛,𝑘(𝑡)𝑖+1,𝑗±𝑑1𝑝=𝑖(1)𝑝+𝑗𝑡𝑖𝑝(𝑖𝑝)!𝑡2𝑑1(𝑝+𝑗)𝑙2𝑑2(𝑝+𝑗)𝑛,𝑘(𝑑1𝑝)!(𝑑1𝑗)!,=𝐜𝑛,𝑘(𝑡)𝑖+1,𝑗±𝑡𝑑𝑗1(𝑑𝑗1)!𝑑1𝑖𝑞=0(𝑡)𝑞𝑡(𝑑1𝑖)𝑝𝑝!((𝑑1𝑖)𝑝)!0.(5.30) As 𝐋𝑛,𝑘 and 𝐑𝑛,𝑘 are constant, we immediately deduce the important relation that, for all 0𝑖𝑑1,(𝝍𝑛,𝑘(𝑡))(𝑖)0,𝑗=(𝝍𝑛,𝑘(𝑡))𝑖,𝑗. This indicates that each finite-dimensional sample paths of our construction has components that satisfies the nondeterministic equation associated with the iteratively integrated Wiener process. Actually, this fact is better stated remembering that the Schauder basis 𝝍𝑛,𝑘 and the corresponding orthonormal basis 𝜙𝑛,𝑘[0,1]1×𝑑 are linked through (3.10), which reads 𝝍𝑛,𝑘0,0𝝍𝑛,𝑘0,𝑑1𝝍𝑛,𝑘𝑑2,0𝝍𝑛,𝑘𝑑2,𝑑1𝝍𝑛,𝑘𝑑1,0𝝍𝑛,𝑘𝑑1,𝑑1=𝝍𝑛,𝑘1,0𝝍𝑛,𝑘1,𝑑1𝝍𝑛,𝑘𝑑1,0𝝍𝑛,𝑘𝑑1,𝑑100+0000𝝍𝑛,𝑘0,0𝝍𝑛,𝑘0,𝑑1.(5.31) Additionally, we realize that the orthonormal basis is entirely determined by the one-dimensional families (𝜙𝑛,𝑘)0,𝑗, which are mutually orthogonal functions satisfying (𝜙𝑛,𝑘)0,𝑗=(𝝍𝑛,𝑘)(𝑑)0,𝑗.

We study in more detail the case of the integrated and doubly-integrated Wiener process (𝑑=2 and 𝑑=3), for which closed-form expressions are provided in Appendices A and B. As expected, the first row of the basis functions for the integrated Wiener process turns out to be the well-known cubic Hermite splines [56]. These functions have been widely used in numerical analysis and actually constitute the basis of the lowest degree in a wider family of bases known as the natural basis of polynomial splines of interpolation [25]. Such bases are used to interpolate data points with constraint of smoothness of different degrees (e.g., the cubic Hermite splines ensure that the resulting interpolation is in 𝐶1[0,1]). The next family of splines of interpolation (corresponding to the 𝐶2 constraint) is naturally retrieved by considering the construction of the doubly-integrated Wiener process: we obtain a family of three 3-dimensional functions that constitutes the columns of a 3×3 matrix that we denote by 𝝍. The top row is made of polynomials of degree five, which have again simple expressions when 𝑚𝑛,𝑘 is the middle of the interval [𝑙𝑛,𝑘,𝑟𝑛,𝑘].

6. Stochastic Calculus from the Hilbert Point of View

Thus far, all calculations, propositions, and theorems are valid for any finite-dimensional the Gauss-Markov process and all the results are valid pathwise, that is, for each sample path. The analysis provides a Hilbert description of the processes as a series of standard Gaussian random variables multiplied by certain specific functions, that form a Schauder basis in the suitable spaces. This new description of Gauss-Markov processes provides a new way for treating problems arising in the study of stochastic processes. As examples of this, we derive the Itô formula and the Girsanov theorem from the Hilbertian viewpoint. Note that these results are equalities in law, that is, dealing with the distribution of stochastic processes, which is a weaker notion compared to the pathwise analysis. In this section, we restrict our analysis to the one-dimensional case for technical simplicity.

The closed-form expressions of the basis of functions 𝜓𝑛,𝑘 in the one-dimensional case are given in Section 5.1. The differential and integral operators associated, introduced in Section 3.1.2 are highly simplified in the one-dimensional case. Let 𝑈 be a bounded open set of [0,1], we denote by 𝐶(𝑈) the space of continuous real functions on 𝑈 and we recall that the topological dual of 𝐶(𝑈) is 𝑅(𝑈), the space of Radon measures on 𝑈. We also introduce 𝐷0(𝑈), the space of test functions in 𝐶(𝑈) that are zero in zero, and whose dual space 𝐷0(𝑈) satisfies 𝐷0(𝑈)𝑅(𝑈).Let 𝑈 be a bounded open neighborhood of [0,1], and denote by 𝐶(𝑈) is the space of continuous real functions on 𝑈, 𝑅(𝑈) its topological dual, the space of Radon measures, 𝐷0(𝑈) the space of test function in 𝐶(𝑈) which are zero at zero and it dual 𝐷0(𝑈)𝑅(𝑈). We consider the Gelfand triple𝐷0(𝑈)𝐶(𝑈)𝐿2(𝑈)𝐷0(𝑈)𝑅(𝑈).(6.1)

The integral operator 𝒦 is defined (and extended by dual pairing) by𝒦[](𝑡)=𝑈𝟙[0,𝑡](𝑠)𝑔𝛼(𝑡)𝑓𝛼(𝑠)𝑑𝑠,(6.2) and the inverse differential operator 𝒟 reads𝒟[](𝑡)=1𝑔𝛼(𝑡)𝑑𝑑𝑡𝑓𝛼(𝑡).(6.3)

Now that we dispose of all the explicit forms of the basis functions and related operators, we are in position to complete our program and start by proving the very important Itô formula and its finite-dimensional counterpart before turning to the Girsanov theorem.

6.1. Itô’s Formula

A very useful theorem in the stochastic processes theory is the Itô formula. We show here that this formula is consistent with the Hilbert framework introduced. Most of the proofs can be found in Appendix E. The proof of Itô formula is based on demonstrating the integration by parts property.

Proposition 6.1 (Integration by parts). Let (𝑋𝑡) and (𝑌𝑡) be two one-dimensional Gauss-Markov processes starting from zero. Then one has the following equality in law: 𝑋𝑡𝑌𝑡=𝑡0𝑋𝑠𝑑𝑌𝑠+𝑡0𝑌𝑠𝑑𝑋𝑠,(6.4) where, for two stochastic processes 𝐴𝑡 and 𝐵𝑡, 𝑡0𝐴𝑠𝑑𝐵𝑠 denotes the Stratonovich integral. In terms of the Itô integral, this formula is written as 𝑋𝑡𝑌𝑡=𝑡0𝑋𝑠𝑑𝑌𝑠+𝑡0𝑌𝑠𝑑𝑋𝑠+𝑋,𝑌𝑡,(6.5) where the brackets denote the mean quadratic variation.

The proof of this proposition is quite technical and is provided in Appendix E. It is based a thorough analysis of the finite-dimensional processes 𝑋𝑁𝑡 and 𝑌𝑁𝑡. For this integration by parts formula and using a density argument, one can recover the more general Itô formula.

Theorem 6.2 (Itô). Let (𝑋𝑡)𝑡 be a Gauss-Markov process and 𝐹𝐶2(). The process 𝑓(𝑋𝑡) is a Markov process and satisfies the following relation in law: 𝑓𝑋𝑡=𝑓𝑋0+𝑡0𝑓𝑋𝑠𝑑𝑋𝑠+12𝑡0𝑓𝑋𝑠𝑑𝑋𝑠.(6.6)

This theorem is proved in Appendix E.

The Itô formula implies in particular that the multiresolution description developed in the paper is valid for every smooth functional of a Gauss-Markov process. In particular, it allows a simple description of exponential functionals of Gaussian Markovian processes, which are of particular interest in mathematics and have many applications, in particular in economics (see, e.g., [57]).

Therefore, we observe that in the view of the paper, Itô formula stems from the nonorthogonal projections of basis element. For multidimensional processes, the proof of the Itô formula is deduced from the one-dimensional proof and would involve the study of the multidimensional bridge formula for 𝐗𝑡 and 𝐘𝑡.

We eventually remark that this section provides us with a finite-dimensional counterpart of the Itô formula for discretized processes, which has important potential applications, and further assesses the suitability of using the finite resolution representation developed in this paper. Indeed, using the framework developed in the present paper allows considering finite-resolution processes and their transformation through nonlinear smooth transformation in a way that is concordant with the standard stochastic calculus processes, since the equation on the transformed process indeed converges towards its Itô representation as the resolution increases.

6.2. Girsanov Formula: A Geometric Viewpoint

In the framework we developed, transforming a process 𝑋 into a process 𝑌 is equivalent to substituting the Schauder construction basis related to 𝑌 for the basis related to 𝑋. Such an operation provides a pathwise mapping for each sample path of 𝑋 onto a sample path of 𝑌 having the same probability density in 𝜉Ω. This fact sheds a new light on the geometry of multidimensional Gauss-Markov processes, since the relationship between two processes is seen as a linear change of basis. In our framework, this relationship between processes is straightforwardly studied in the finite rank approximations of the processes up to a certain resolution. Technical intricacy is nevertheless raised when dealing with the representation of the process itself in the infinite-dimensional Hilbert spaces. We solve these technical issues here and show that in the limit 𝑁 one recovers Girsanov theorem as a limit of the linear transformations between the Gauss-Markov processes.

The general problem consists therefore in studying the relationship between two real Gauss-Markov processes 𝑋 and 𝑌 that are defined by𝑑𝑋𝑡=𝛼𝑋(𝑡)𝑋𝑡𝑑𝑡+Γ𝑋(𝑡)𝑑𝑊𝑡,𝑑𝑌𝑡=𝛼𝑌(𝑡)𝑋𝑡𝑑𝑡+Γ𝑌(𝑡)𝑑𝑊𝑡.(6.7)

We have noticed that the spaces 𝑥Ω are the same in the one-dimensional case as long as both Γ𝑋 and Γ𝑌 never vanish and therefore make this assumption here. In order to further simplify the problem, we assume that 𝛾𝑋,𝑌=Γ𝑋/Γ𝑌 is continuously differentiable. This assumption allows us to introduce the process 𝑍𝑡=𝛾𝑋,𝑌(𝑡)𝑌𝑡 that satisfies the stochastic differential equation𝑑𝑍𝑡=𝑑𝑑𝑡𝛾𝑋,𝑌(𝑡)𝑌𝑡𝑑𝑡+𝛾𝑋,𝑌(𝑡)𝑑𝑌𝑡=𝑑𝑑𝑡𝛾𝑋,𝑌(𝑡)+𝛾𝑋,𝑌(𝑡)𝛼𝑌(𝑡)𝑌𝑡𝑑𝑡+Γ𝑋(𝑡)𝑑𝑊𝑡=𝛼𝑍(𝑡)𝑍𝑡𝑑𝑡+Γ𝑋(𝑡)𝑑𝑊𝑡,(6.8) with 𝛼𝑍(𝑡)=(𝑑/𝑑𝑡)(𝛾𝑋,𝑌(𝑡))𝛾𝑋,𝑌(𝑡)1+𝛼𝑌(𝑡). Moreover, if 𝑍𝜓𝑛,𝑘 and 𝑌𝜓𝑛,𝑘 are the bases of functions that describe the process 𝑍 and 𝑌, respectively, we have 𝑍𝜓𝑛,𝑘=𝛾𝑋,𝑌𝑌𝜓𝑛,𝑘.

The previous remarks allow us to restrict without loss of generality our study to the processes defined for same function Γ, thus reducing the parameterization of the Gauss-Markov processes to the linear coefficient 𝛼. Observe that in the classical stochastic calculus theory, it is well known that such hypotheses are necessary for the process 𝑋 to be absolutely continuous with respect to 𝑌 (through the use of the Girsanov theorem).

Let us now consider that 𝛼,𝛽, and Γ three real Hölder continuous real functions, and introduce 𝛼𝑋 and 𝛽𝑋 solutions of the equations𝑑𝛼𝑋𝑡=𝛼(𝑡)𝛼𝑋𝑡𝑑𝑡+Γ(𝑡)𝑑𝑊𝑡,𝑑𝛽𝑋𝑡=𝛽(𝑡)𝛽𝑋𝑡𝑑𝑡+Γ(𝑡)𝑑𝑊𝑡.(6.9) All the functions and tools related to the process 𝛼𝑋 (resp., 𝛽𝑋) will be indexed by 𝛼 (𝛽) in the sequel.

6.2.1. Lift Operators

Depending on the space we are considering (either coefficients or trajectories), we define the two following operators mapping the process 𝛼𝑋𝑡 on 𝛽𝑋𝑡.(1)The coefficients lift operator 𝛼,𝛽𝐺 is the linear operator mapping in 𝜉Ω the process 𝛼𝑋 on the process 𝛽𝑋: 𝛼,𝛽𝐺=𝛽Δ𝛼Ψ𝜉Ω,𝜉Ω𝜉Ω,𝜉Ω.(6.10) For any 𝜉𝜉Ω, the operator 𝛼,𝛽𝐺 maps a sample path of 𝛼𝑋 on a sample path of 𝛽𝑋.(2)The process lift operator 𝛼,𝛽𝐹 is the linear operator mapping in 𝑥Ω the process 𝛼𝑋 on the process 𝛽𝑋: 𝛼,𝛽𝐻=𝛼Ψ𝛼Δ𝑥Ω,𝑥Ω𝑥Ω,𝑥Ω.(6.11) We summarize the properties of these operators now.

Proposition 6.3. The operators 𝛼,𝛽𝐺 and 𝛼,𝛽𝐻 satisfy the following properties.(i) They are linear measurable bijections(ii)For every 𝑁>0, the function 𝛼,𝛽𝐺𝑁=𝑃𝑁𝛼,𝛽𝐺𝐼𝑁𝜉Ω𝑁𝜉Ω𝑁 (resp., 𝛼,𝛽𝐻𝑁=𝑃𝑁𝛼,𝛽𝐻𝐼𝑁𝜉Ω𝑁𝜉Ω𝑁) is a finite-dimensional linear operator, whose matrix representation is triangular in the natural basis of 𝜉Ω𝑁 (resp., 𝜉Ω𝑁) and whose eigenvalues 𝛼,𝛽𝜈𝑛,𝑘 are given by 𝛼,𝛽𝜈𝑛,𝑘=𝑔𝛼𝑚𝑛,𝑘𝑔𝛽𝑚𝑛,𝑘𝛽𝑀𝑛,𝑘𝛼𝑀𝑛,𝑘,0𝑛𝑁,0𝑘<2𝑁1(6.12) (resp., 𝛽,𝛼𝜈𝑛,𝑘=(𝛼,𝛽𝜈𝑛,𝑘)1).(iii)𝛼,𝛽𝐺 and 𝛼,𝛽𝐻 are bounded operators for the spectral norm with 𝛼,𝛽𝐺2=sup𝑛sup𝑘𝛼,𝛽𝜈𝑛,𝑘sup𝑔𝛼inf𝑔𝛽sup𝑓2𝛼inf𝑓2𝛽<,(6.13) and 𝛼,𝛽𝐻2=𝛽,𝛼𝐺2<.(iv)The determinants of 𝛼,𝛽𝐺𝑁 (denoted by 𝛼,𝛽𝐽𝑁) and 𝛼,𝛽𝐻𝑁 admit a limit when 𝑁 tends to infinity: 𝛼,𝛽𝐽=lim𝑁𝛼,𝛽𝐽𝑁=exp1210(𝛼(𝑡)𝛽(𝑡))𝑑𝑡,lim𝑁det𝛼,𝛽𝐻𝑁=exp1210(𝛽(𝑡)𝛼(𝑡))𝑑𝑡=𝛽,𝛼𝐽.(6.14)

The proof of these properties elementary stems from the analysis done on the functions Ψ and Δ that were previously performed, and these are detailed in Appendix C.

6.2.2. Radon-Nikodym Derivatives

From the properties proved on the lift operators, we are in position to further analyze the relationship between the probability distributions of 𝛼𝑋 and 𝛽𝑋. We first consider the finite-dimensional processes 𝛼𝑋𝑁 and 𝛽𝑋𝑁. We emphasize that, throughout this section, all equalities are true pathwise.

Lemma 6.4. Given the finite-dimensional measures 𝑃𝑁𝛼 and 𝑃𝑁𝛽, the Radon-Nikodym derivative of 𝑃𝑁𝛽 with respect to 𝑃𝑁𝛼 satisfies 𝑑𝑃𝑁𝛽𝑑𝑃𝑁𝛼(𝜔)=𝛼,𝛽𝐽𝑁exp12Ξ𝑁(𝜔)𝑇𝛼,𝛽𝑆𝑁𝐼𝑑𝜉Ω𝑁Ξ𝑁(𝜔)(6.15) with 𝛼,𝛽𝑆𝑁=𝛼,𝛽𝐺𝑁𝑇𝛼,𝛽𝐺𝑁 and the equality is true pathwise.

Proof. In the finite-dimensional case, for all 𝑁>0, we can write that 𝑃𝑁𝛼, 𝑃𝑁𝛽𝑝𝑁𝛼, and the Lebesgue measure 𝑥Ω𝑁 are mutually absolutely continuous: we denote by 𝑝𝑁𝛼 and 𝑝𝑁𝛽 the Gaussian density of 𝑃𝑁𝛼 and 𝑃𝑁𝛽 with respect to the Lebesgue measure on 𝑥Ω𝑁. Therefore, in the finite-dimensional case, the Radon-Nikodym derivative of 𝑃𝑁𝛽 with respect to 𝑃𝑁𝛼 is defined to be pathwise and is simply given by the quotient of the density of the vector {𝛽𝑋𝑁(𝑚𝑖,𝑗)} with the density of the vector {𝛼𝑋𝑁(𝑚𝑖,𝑗)} for 0𝑖𝑁, 0𝑗<2𝑖1, that is, 𝑑𝑃𝑁𝛽𝑑𝑃𝑁𝛼(𝜔)=𝑝𝑁𝛽𝛼𝑋𝑁(𝜔)𝑝𝑁𝛼𝛼𝑋𝑁(𝜔)=det𝛼Σ𝑁det𝛽Σ𝑁exp12𝛼𝑋𝑁(𝜔)𝑇𝛽Σ1𝑁𝛼Σ1𝑁𝛼𝑋𝑁(𝜔).(6.16) We first make explicit det𝛼Σ𝑁det𝛽Σ𝑁=det𝛼Σ𝑁𝛽Σ1𝑁=det𝛼Ψ𝑁𝛼Ψ𝑁𝑇𝛽Δ𝑁𝑇𝛽Δ𝑁=det𝛽Δ𝑁𝛼Ψ𝑁𝛼Ψ𝑁𝑇𝛽Δ𝑁𝑇=det𝛼,𝛽𝐺𝑁𝛼,𝛽𝐺𝑁𝑇=det𝛼,𝛽𝐺𝑁2.(6.17) Then, we rearrange the exponent using the Cholesky decomposition: 𝛽Σ1𝑁𝛼Σ1𝑁=𝛽Δ𝑁𝑇𝛽Δ𝑁𝛼Δ𝑁𝑇𝛼Δ𝑁,(6.18) so that we write the exponent of (6.16) as 𝛼𝑋𝑁𝑇𝛽Δ𝑁𝑇𝛽Δ𝑁𝛼Δ𝑁𝑇𝛼Δ𝑁𝛼𝑋𝑁=Ξ𝑁𝑇𝛼Ψ𝑁𝑇𝛽Δ𝑁𝑇𝛽Δ𝑁𝛼Δ𝑁𝑇𝛼Δ𝑁𝛼Ψ𝑁(𝜔)Ξ𝑁=Ξ𝑁𝑇𝛼,𝛽𝐺𝑁𝑇𝛼,𝛽𝐺𝑁𝐼𝑑𝜉Ω𝑁Ξ𝑁.(6.19) We finally reformulate (6.16) as 𝑑𝑃𝑁𝛽𝑑𝑃𝑁𝛼(𝜔)=𝛼,𝛽𝐽𝑁exp12Ξ𝑁(𝜔)𝑇𝛼,𝛽𝐺𝑁𝑇𝛼,𝛽𝐺𝑁𝐼𝑑𝜉Ω𝑁Ξ𝑁(𝜔).(6.20)

Let us now justify from a geometrical point of view why this formula is a direct consequence of the finite-dimensional change of variable formula on the model space 𝜉Ω𝑁. If we introduce 𝛼Δ𝑁, the coefficient application related to 𝛼𝑋𝑁, we know that Ξ𝑁=𝛼Δ𝑁(𝛼𝑋𝑁) follows a normal law 𝒩(𝟎,𝐈𝜉Ω𝑁). We denote by 𝑝𝑁𝜉 its standard Gaussian density with respect to the Lebesgue measure on 𝜉Ω𝑁. We also know that𝛼Δ𝑁𝛽𝑋𝑁=𝛼Δ𝑁𝛼,𝛽𝐹𝑁𝛼𝑋𝑁=𝛽,𝛼𝐺𝑁𝛼Δ𝑁𝛼𝑋𝑁=𝛽,𝛼𝐺𝑁Ξ𝑁.(6.21)

Since 𝛽,𝛼𝐺𝑁 is linear, the change of the variable formula directly entails that 𝛽,𝛼𝐺𝑁(Ξ𝑁) admits on 𝜉Ω𝑁𝑝𝑁𝛽,𝛼𝜉𝑁=||det𝛼,𝛽𝐺||𝑝𝑁𝜉𝜉𝑁𝑇𝛼,𝛽𝐺𝑁𝑇𝛼,𝛽𝐺𝑁𝜉𝑁(6.22) as density with respect to the Lebesgue measure. Consider now 𝐵 as a measurable set of (𝑥Ω𝑁,(𝑥Ω𝑁)); then we have𝑃𝑁𝛽(𝐵)=𝛼Δ𝑁(𝐵)𝑝𝑁𝛽,𝛼𝜉𝑁𝑑𝜉𝑁=𝛼Δ𝑁(𝐵)𝑝𝑁𝛽,𝛼𝜉𝑁𝑝𝑁𝜉𝜉𝑁𝑝𝑁𝜉𝜉𝑁𝑑𝜉𝑁=𝐵𝑝𝑁𝛽,𝛼𝛼Δ𝑁𝑋𝑁𝑝𝑁𝜉𝛼Δ𝑁𝑋𝑁𝑑𝑃𝑁𝛼𝑋𝑁,(6.23) from which we immediately conclude.

6.2.3. The Trace Class Operator

The pathwise expression of the Radon-Nikodym derivative extends to the infinite-dimensional representation of 𝛼𝑋 and 𝛽𝑋. This extension involves technical analysis on the infinite-dimensional Hilbert space 𝑙2(). We have shown in Proposition 6.3 that the application 𝛼,𝛽𝐺 was bounded for the spectral norm. Therefore, we have, for any 𝜉 in 𝑙2(), the inequality𝛼,𝛽𝐺(𝜉)2𝛼,𝛽𝐺2𝜉2(6.24) implying that 𝛼,𝛽𝐺 maps 𝑙2() into 𝑙2(). We can then define the adjoint operator 𝛼,𝛽𝐺𝑇 from 𝑙2() to 𝑙2(), which is given by𝜉,𝜂𝑙2(),𝛼,𝛽𝐺𝑇(𝜂),𝜉=𝜂,𝛼,𝛽𝐺(𝜉).(6.25) Let us now consider the self-adjoint operator 𝛼,𝛽𝑆=𝛼,𝛽𝐺𝑇𝛼,𝛽𝐺𝑙2()𝑙2(). This operator is the infinite-dimensional counterpart of the matrix 𝛼,𝛽𝑆𝑁.

Lemma 6.5. Considering the coefficients of the matrix representation of 𝛼,𝛽𝑆 in the natural basis 𝑒𝑛,𝑘 of 𝑙2() given as 𝛼,𝛽𝑆𝑛,𝑘𝑝,𝑞=(𝑒𝑛,𝑘,𝛼,𝛽𝑆(𝑒𝑝,𝑞)), one has 𝛼,𝛽𝑆𝑛,𝑘𝑝,𝑞=10𝛼𝜙𝑛,𝑘(𝑡)+(𝛼(𝑡)𝛽(𝑡))Γ(𝑡)𝛼𝜓𝑛,𝑘(𝑡)𝛼𝜙𝑝,𝑞(𝑡)+(𝛼(𝑡)𝛽(𝑡))Γ(𝑡)𝛼𝜓𝑝,𝑞(𝑡)𝑑𝑡.(6.26)

Proof. Assume that max(𝑛,𝑝)𝑁 and that (𝑛,𝑘) and (𝑝,𝑞)𝑁. With the notations used previously with 𝑈, an open neighbourhood of [0,1], we have 𝛼,𝛽𝑆𝑛,𝑘𝑝,𝑞=𝛼,𝛽𝐺𝑒𝑛,𝑘,𝛼,𝛽𝐺𝑒𝑝,𝑞=(𝑖,𝑗)𝑈𝛽𝛿𝑖,𝑗(𝑡)𝛼𝜓𝑛,𝑘(𝑡)𝑑𝑡𝑈𝛽𝛿𝑖,𝑗(𝑡)𝛼𝜓𝑝,𝑞(𝑡)𝑑𝑡.(6.27) Since we have by definition 𝛼𝜓𝑛,𝑘=𝛼𝒦[𝛼𝜙𝑛,𝑘], 𝛽𝛿𝑖,𝑗=𝛽𝒟[𝛽𝜙i,𝑗], 𝛼𝒦1=𝛼𝒟, 𝛽𝒟1=𝛽𝒦 on the space 𝐷0(𝑈), we have 𝛽𝜙𝑖,𝑗(𝑡)=𝛽𝒦𝛽𝛿𝑖,𝑗(𝑡)=𝑓𝛽(𝑡)𝑈𝑔𝛽(𝑠)𝛽𝛿𝑖,𝑗(𝑠)𝑑𝑠,(6.28)𝛼𝜙𝑛,𝑘(𝑡)=𝛼𝒟𝛼𝜓𝑛,𝑘(𝑡)=1𝑓𝛼(𝑡)𝑑𝑑𝑡𝛼𝜓𝑛,𝑘(𝑡)𝑔𝛼(𝑡).(6.29) From there using (6.28), we can write the integration by part formula in the sense of the generalized functions to get 𝑈𝛽𝛿𝑖,𝑗(𝑡)𝛼𝜓𝑛,𝑘(𝑡)𝑑𝑡=𝑈1𝑓𝛽(𝑡)𝑑𝑑𝑡𝛼𝜓𝑛,𝑘(𝑡)𝑔𝛽(𝑡)𝛽𝜙𝑖,𝑗(𝑡)𝑑𝑡=𝑈𝛽𝒟𝛼𝜓𝑛,𝑘(𝑡)𝛽𝜙𝑖,𝑗(𝑡)𝑑𝑡.(6.30) We now compute using (6.29): 𝛽𝒟𝛼𝜓𝑛,𝑘(𝑡)=1𝑓𝛽(𝑡)𝑑𝑑𝑡𝛼𝜓𝑛,𝑘(𝑡)𝑔𝛼(𝑡)𝑔𝛼(𝑡)𝑔𝛽(𝑡)=𝑔𝛼(𝑡)𝑓(𝑡)𝑔𝛽(𝑡)𝑓𝛽(𝑡)𝛼𝜙(𝑡)𝑑𝑑𝑡𝑔𝛼(𝑡)𝑔𝛽(𝑡)𝛼𝜓𝑛,𝑘(𝑡)𝑔𝛽(𝑡)𝑓𝛽(𝑡).(6.31) Specifying 𝑔𝛼, 𝑔𝛽, 𝑓𝛼, and 𝑓𝛽 and recalling the relations 𝑔𝛼(𝑡)𝑓𝛼(𝑡)=𝑔𝛽(𝑡)𝑓𝛽(𝑡)=Γ(𝑡),𝑑𝑑𝑡𝑔𝛼(𝑡)𝑔𝛽(𝑡)=(𝛼(𝑡)𝛽(𝑡))𝑔𝛼(𝑡)𝑔𝛽(𝑡),(6.32) we rewrite the function (6.31) in 𝐿2[0,1] as 𝛽𝒟𝛼𝜓𝑛,𝑘(𝑡)=1𝑓𝛽(𝑡)𝑑𝑑𝑡𝛼𝜓𝑛,𝑘(𝑡)𝑔𝛽(𝑡)=𝛼𝜙𝑛,𝑘(𝑡)+(𝛼(𝑡)𝛽(𝑡))Γ(𝑡)𝛼𝜓𝑛,𝑘(𝑡).(6.33) Now, since the family 𝛽𝜙𝑛,𝑘 forms a complete orthonormal system of 𝐿2[0,1] and is zero outside [0,1], by the Parseval identity, expression (6.27) can be written as the scalar product: 𝛼,𝛽𝑆𝑛,𝑘𝑝,𝑞=10𝛼𝜙𝑛,𝑘(𝑡)+(𝛼(𝑡)𝛽(𝑡))Γ(𝑡)𝛼𝜓𝑛,𝑘(𝑡)𝛼𝜙𝑝,𝑞(𝑡)+(𝛼(𝑡)𝛽(𝑡))Γ(𝑡)𝛼𝜓𝑝,𝑞(𝑡)𝑑𝑡,(6.34) which ends the proof of the lemma.

Notice that we can further simplify expression (6.26):𝛼,𝛽𝑆𝑛,𝑘𝑝,𝑞=𝛿𝑛,𝑘𝑝,𝑞+10(𝛼(𝑡)𝛽(𝑡))2Γ(𝑡)𝛼𝜓𝑛,𝑘(𝑡)𝛼𝜓𝑝,𝑞(𝑡)𝑑𝑡+10(𝛼(𝑡)𝛽(𝑡))Γ(𝑡)𝛼𝜙𝑛,𝑘(𝑡)𝛼𝜓𝑝,𝑞(𝑡)+𝛼𝜙𝑝,𝑞(𝑡)𝛼𝜓𝑛,𝑘(𝑡)𝑑𝑡.(6.35) We are now in a position to show that the operator 𝛼,𝛽𝑆𝐼𝑑 can be seen as the limit of the finite-dimensional operator 𝛼,𝛽𝑆𝑁𝐼𝑑𝜉Ω𝑁, in the following sense.

Theorem 6.6. The operator 𝛼,𝛽𝑆𝐼𝑑𝑙2()𝑙2() is a trace class operator, whose trace is given by Tr𝛼,𝛽𝑆𝐼𝑑=10(𝛼(𝑡)𝛽(𝑡))𝑑𝑡+10𝛼(𝑡)𝑓𝛼(𝑡)2(𝛼(𝑡)𝛽(𝑡))2𝑑𝑡.(6.36)

We prove this essential point in Appendix F. The proof consists in showing that the operator 𝛼,𝛽𝑆𝐼𝑑 is isometric to a Hilbert-Schmidt operator whose trace can be computed straightforwardly.

6.2.4. The Girsanov Theorem

We now proceed to prove the Girsanov theorem by extending the domain of the quadratic form associated with 𝛼,𝛽𝑆𝐼𝑑 to the space 𝜉Ω, which can only be done in law.

Theorem 6.7. In the infinite-dimensional case, the Radon-Nikodym derivative of 𝑃𝛽=𝑃𝛽𝑋1 with respect to 𝑃𝛼=𝑃𝛼𝑋1 reads 𝑑𝑃𝛽(𝜔)𝑑𝑃𝛼(𝜔)=exp1210𝛽(𝑡)𝛼(𝑡)𝑑𝑡+Ξ(𝜔)𝑇𝛼,𝛽𝑆𝐼𝑑𝜉ΩΞ(𝜔),(6.37) which in terms of the Itô stochastic integral reads 𝑑𝑃𝛽(𝜔)𝑑𝑃𝛼(𝜔)=exp10𝛽(𝑡)𝛼(𝑡)𝑓𝛼2(𝑡)𝛼𝑋𝑡(𝜔)𝑔𝛼(𝑡)𝑑𝛼𝑋𝑡(𝜔)𝑔𝛼(𝑡)1210(𝛽(𝑡)𝛼(𝑡))2𝑓𝛼2(𝑡)𝛼𝑋𝑡(𝜔)𝑔𝛼(𝑡)2𝑑𝑡.(6.38)

In order to demonstrate the Girsanov theorem from our geometrical point of view, we need to establish the following result.

Lemma 6.8. The positive definite quadratic form on 𝑙2()×𝑙2() associated with operator 𝛼,𝛽𝑆𝐼𝑑𝑙2()𝑙2() is well defined on 𝜉Ω. Moreover, for all 𝜉Ω, 𝜉,𝛼,𝛽𝑆𝐼𝑑𝜉Ω(𝜉)=210𝛼(𝑡)𝛽(𝑡)𝑓𝛼2(𝑡)𝛼𝑋𝑡(𝜉)𝑔𝛼(𝑡)𝑑𝛼𝑋𝑡(𝜉)𝑔𝛼(𝑡)+10(𝛼(𝑡)𝛽(𝑡))2𝑓𝛼2(𝑡)𝛼𝑋𝑡(𝜉)𝑔𝛼(𝑡)2𝑑𝑡,(6.39) where 𝛼𝑋𝑡(𝜉)=𝛼Φ(𝜉) and refers to the Stratonovich integral and the equality is true in law.

Proof of Theorem 6.7. We start by writing the finite-dimensional Radon-Nikodym derivative 𝑑𝑃𝛽(𝜔)𝑑𝑃𝛼(𝜔)=𝛼,𝛽𝐽𝑁lim𝑁exp12Ξ𝑁(𝜔)𝑇𝛼,𝛽𝑆𝑁𝐼𝑑𝜉Ω𝑁Ξ𝑁(𝜔).(6.40) By Proposition 6.3, we have 𝛼,𝛽𝐽=lim𝑁𝛼,𝛽𝐽𝑁=1210(𝛼(𝑡)𝛽(𝑡))𝑑𝑡.(6.41) If, as usual, Ξ denotes a recursively indexed infinite-dimensional vector of independent variables with law 𝒩(0,1) and Ξ𝑁=𝜉𝑃𝑁Ξ, writing 𝜉𝑛,𝑘=Ξ𝑛,𝑘(𝜔), we have Ξ𝑁(𝜔)𝑇𝛼,𝛽𝑆𝑁𝐼𝑑𝜉Ω𝑁Ξ𝑁(𝜔)=𝑁𝑛=0𝑁𝑝=00𝑘<2𝑛10𝑞<2𝑝1𝜉𝑛,𝑘𝛼,𝛽𝑆𝑁𝐼𝑑𝜉Ω𝑁𝑛,𝑘𝑝,𝑞𝜉𝑝,𝑞.(6.42) We know that 𝜉 is almost surely in 𝜉Ω, and, by Lemma 6.8, we also know that on 𝜉Ω×𝜉Ωlim𝑁𝜉,𝛼,𝛽𝑆𝑁𝐼𝑑𝜉Ω𝑁(𝜉)=𝜉,𝛼,𝛽𝑆𝐼𝑑𝜉Ω(𝜉),(6.43) so that we can effectively write the infinite-dimensional Radon-Nikodym derivative on 𝜉Ω as the point-wise limit of the finite-dimensional one on 𝜉Ω𝑁 through the projectors 𝜉𝑃𝑁: 𝑑𝑃𝛽𝑑𝑃𝛼(𝜔)=lim𝑁𝑑𝑃𝑁𝛽𝑑𝑃𝑁𝛼(𝜔),(6.44) which directly yields formula (6.37).
The derivation of the Girsanov formula (6.40) from (6.37) comes from the relationship between the Stratonovich and Itô formulas since the quadratic variation of 𝛼𝑋𝑡/𝑔𝛼(𝑡) and (𝛼(𝑡)𝛽(𝑡))/𝑓2𝛼(𝑡)×𝛼𝑋t/𝑔𝛼(𝑡) reads 10𝐸𝑡0𝑓𝛼(𝑠)𝑑𝑊𝑠,𝛼(𝑡)𝛽(𝑡)𝑓𝛼2(𝑡)𝑡0𝑓𝛼(𝑠)𝑑𝑊𝑠=10(𝛼(𝑡)𝛽(𝑡))𝑑𝑡.(6.45) Therefore, the expression of the Radon-Nikodym derivative in Lemma 6.8 can be written in terms of the Itô integrals as 𝜉,𝛼,𝛽𝑆𝐼𝑑𝜉Ω(𝜉)=10(𝛼(𝑡)𝛽(𝑡))𝑑𝑡+210𝛼(𝑡)𝛽(𝑡)𝑓2(𝑡)𝛼𝑋𝑡(𝜉)𝑔𝛼(𝑡)𝑑𝛼𝑋𝑡(𝜉)𝑔𝛼(𝑡)+10(𝛼(𝑡)𝛽(𝑡))2𝑓2(𝑡)𝛼𝑋𝑡(𝜉)𝑔𝛼(𝑡)2𝑑𝑡(6.46) and the Radon-Nikodym derivative as 𝑑𝑃𝛽𝑑𝑃𝛼(𝜔)=exp10𝛽(𝑡)𝛼(𝑡)𝑓𝛼2(𝑡)𝛼𝑋𝑡(𝜔)𝑔𝛼(𝑡)𝑑𝛼𝑋𝑡(𝜔)𝑔𝛼(𝑡)1210(𝛽(𝑡)𝛼(𝑡))2𝑓𝛼2(𝑡)𝛼𝑋𝑡(𝜔)𝑔𝛼(𝑡)2𝑑𝑡.(6.47)

Observe that, if 𝛼(𝑡)=0, we recover the familiar expression𝑑𝑃𝛽𝑑𝑃𝛼(𝜔)=exp10𝛽(𝑡)Γ(𝑡)𝑊𝑡(𝜔)𝑑𝑊𝑡(𝜔)1210𝛽(𝑡)2Γ(𝑡)𝑊(𝜔)2𝑡𝑑𝑡.(6.48)

Conclusion and Perspectives
The discrete construction we present displays both analytical and numerical interests for further applications. From the analysis viewpoint, even if the basis does not exhibit the same orthogonal properties as the Karhunen-Loève decomposition, it has the important advantage of saving the structure of sample paths through its property of strong pathwise convergence and of providing a multiscale representation of the processes, which contrasts with the convergence in the mean of the Karhunen-Loève decomposition. From the numerical viewpoint, three Haar-like properties make our decomposition particularly suitable for certain numerical computations: (i) all basis elements have compact support on an open interval that has the structure of dyadic rational endpoints, (ii) these intervals are nested and become smaller for larger indices of the basis element, and (iii) for any interval endpoint, only a finite number of basis elements are nonzero at that point. Thus the expansion in our basis, when evaluated at an interval endpoint (e.g., dyadic rational), terminates in a finite number of steps. Moreover, the very nature of the construction based on an increasingly refined description of the sample paths paves the way to coarse-graining approaches similar to wavelet decompositions in signal processing. In view of this, our framework offers promising applications.Dichotomic Search of First-Hitting Times
The first application we envisage concerns the problem of first-hitting times. Because of its manifold applications, finding the time when a process first exits a given region is a central question of stochastic calculus. However, closed-form theoretical results are scarce and one often has to resort to numerical algorithms [59]. In this regard, the multiresolution property suggests an exact scheme to simulate sample paths of a Gaussian Markov process 𝐗 in an iterative “top-down’’ fashion. Assuming the intervals are dyadic rational and that we have a conditional knowledge of a sample path on the dyadic points of 𝐷𝑁={𝑘2𝑁0𝑘2𝑁}, one can decide to further the simulation of this sample path at any time 𝑡 in 𝐷𝑁+1 by drawing a point according to the conditional law of 𝐗𝑡 given {𝐗𝑡}𝑡𝐷𝑁, which is simply expressed in the framework of our construction. This property can be used for great advantages in numerical computations such as dichotomic search algorithms for first passage times: the key element is to find an estimate of the true conditional probability that a hitting time has occurred when knowing the value of the process at two given times, one in the past and one in the future. With such an estimate, an efficient strategy to look for passage times consists in refining the sample path when and only when its trajectory is estimated likely to actually cross the barrier. Thus the sample path of the process is represented at poor temporal resolution when it is far from the boundary and at increasingly higher resolution closer to the boundary. Such an algorithmic principle achieves a high level of precision in the computation of the first-hitting time, while demanding far less operation than usual stochastic Runge-Kutta scheme. This approach has been successfully implemented for the one-dimensional case [58], see Figure 4. In that article, the precision of the algorithm is controlled as well as the probability to evaluate a first hitting time substantially different from the actual value. The approach proves to be extremely efficient compared to customary methods. The general multidimensional approach proposed in the present paper allows direct generalization of these results to the computation of exit times in any dimension and for general smooth sets [3032].
Gaussian Deformation Modes in Nonlinear Diffusions
The present study is developed for the Gauss-Markov systems. However, many models arising in applied science present nonlinearities, and in that case, the construction based on a sum of Gaussian random variables will not generalize. However, the Gaussian case treated here can nevertheless be applied to perturbation of nonlinear differential equations with small noise. Let 𝐅×𝑑𝑑 be a nonlinear time-varying vector field, and let us assume that 𝐗0(𝑡) is a stable (attractive) solution of the dynamical system: 𝑑𝐗𝑑𝑡=𝐅(𝐭,𝐗).(6.49) This function 𝐗0 can for instance be a fixed point (in which case it is a constant), a cycle (in which case it is periodic), or a general attractive orbit of the system. In the deterministic case, any solution having its initial condition in a given neighbourhood in ×𝑑 of the solution will asymptotically converge towards the solution, and therefore perturbations of the solutions are bounded. Let us now consider that the system is subject to a small amount of noise and define 𝐘𝑑 as the solution of the stochastic nonlinear differential equation: 𝑑𝐘𝑡=𝐅𝑡,𝐘𝑡𝑑𝑡+𝜀𝚪𝑡,𝑌𝑡𝑑𝐖𝑡.(6.50) Assuming that the noise is small (i.e., 𝜀 is a small parameter), because of the attractivity of the solution 𝐗0(𝑡), the function 𝐘(𝑡) will remain very close to 𝐗0(𝑡) (at least in a bounded time interval). In this region, we define 𝜀𝐙𝑡=𝐘𝑡𝐗0(𝑡). This stochastic variable is the solution of the equation 𝑑𝐙𝑡=1𝜀𝐅𝑡,𝐗0(𝑡)+𝜀𝐙𝑡𝐅𝑡,𝐗0(𝑡)+𝜀𝚪𝑡,𝐗0(𝑡)+𝜀𝐙𝑡𝑑𝐖𝑡=𝑥𝐅𝑡,𝐗0(𝑡)𝐙𝑡+𝚪𝑡,𝐗0(𝑡)𝑑𝐖𝑡+𝑂(𝜀).(6.51) The solution at the first order in 𝜀 is therefore the multidimensional Gaussian process with nonconstant coefficients: 𝑑𝐙𝑡=𝑥𝐅𝑡,𝐗0(𝑡)𝐙𝑡+𝚪𝑡,𝐗0(𝑡)𝑑𝐖𝑡,(6.52) and our theory describes the solutions in a multiresolution framework. Notice that, in that perspective, our basis functions 𝝍𝑛,𝑘 can be seen as increasingly finer modes of deformation of a deterministic trajectory. This approach appears particularly relevant to the theory of weakly interconnected neural oscillators in computational neuroscience [60]. Indeed, one of the most popular approaches of this field, the phase model theory, formally consists in studying how perturbations are integrated in the neighborhood of an attracting cycle [61, 62].
All these instances are exemplary of how our multiresolution description of the Gauss-Markov processes offers a simple yet rigorous tool to broach a large number of open problems and promises fascinating applications both in theoretical and in applied science.

Appendices

A. Formulae of the Basis for the Integrated Wiener Process

In the case of the primitive of the Wiener process, straightforward linear algebra computations lead to the two bases of functions ((𝜓𝑛,𝑘)1,1,(𝜓𝑛,𝑘)2,1) and ((𝜓𝑛,𝑘)1,2,(𝜓𝑛,𝑘)2,2) having the following expressions𝜓𝑛,𝑘1,1(𝑡)=𝜎𝑛,𝑘1,1𝑡𝑙𝑛,𝑘𝑚𝑛,𝑘𝑙𝑛,𝑘21+2𝑚𝑛,𝑘𝑡𝑚𝑛,𝑘𝑙𝑛,𝑘,𝑙𝑛,𝑘𝑡𝑚𝑛,𝑘,𝜎𝑛,𝑘1,1𝑟𝑛,𝑘𝑡𝑟𝑛,𝑘𝑚𝑛,𝑘21+2𝑡𝑚𝑛,𝑘𝑟𝑛,𝑘𝑚𝑛,𝑘,𝑚𝑛,𝑘𝑡𝑟𝑛,𝑘,𝜓𝑛,𝑘2,1(𝑡)=𝜎𝑛,𝑘1,16𝑡𝑙𝑛,𝑘𝑚𝑛,𝑘𝑡𝑚𝑛,𝑘𝑙𝑛,𝑘3,𝑙𝑛,𝑘𝑡𝑚𝑛,𝑘,𝜎𝑛,𝑘1,16𝑟𝑛,𝑘𝑡𝑡𝑚𝑛,𝑘𝑟𝑛,𝑘𝑚𝑛,𝑘3,𝑚𝑛,𝑘𝑡𝑟𝑛,𝑘,𝜓𝑛,𝑘1,2(𝑡)=𝜎𝑛,𝑘2,2𝑚𝑛,𝑘𝑡𝑡𝑙𝑛,𝑘𝑚𝑛,𝑘𝑙𝑛,𝑘2,𝑙𝑛,𝑘𝑡𝑚𝑛,𝑘,𝜎𝑛,𝑘2,2𝑡𝑚𝑛,𝑘𝑟𝑛,𝑘𝑡𝑟𝑛,𝑘𝑚𝑛,𝑘2,𝑚𝑛,𝑘𝑡𝑟𝑛,𝑘,𝜓𝑛,𝑘2,2(𝑡)=𝜎𝑛,𝑘2,2𝑡𝑙𝑛,𝑘𝑚𝑛,𝑘𝑙𝑛,𝑘212𝑚𝑛,𝑘𝑡𝑡𝑙𝑛,𝑘,𝑙𝑛,𝑘𝑡𝑚𝑛,𝑘,𝜎𝑛,𝑘2,2𝑟𝑛,𝑘𝑡𝑟𝑛,𝑘𝑚𝑛,𝑘212𝑡𝑚𝑛,𝑘𝑟𝑛,𝑘𝑡,𝑚𝑛,𝑘𝑡𝑟𝑛,𝑘,(A.1) where𝜎𝑛,𝑘1,1=1196𝑟𝑛,𝑘𝑙𝑛,𝑘3,𝜎𝑛,𝑘2,2=132𝑟𝑛,𝑘𝑙𝑛,𝑘(A.2) are the diagonal components of the (diagonal) matrix 𝜎𝑛,𝑘. As expected, we notice that the differential structure of the process is conserved at any finite rank since we have𝑑𝑑𝑡𝜓𝑛,𝑘1,1(𝑡)=𝜓𝑛,𝑘2,1(𝑡),𝑑𝑑𝑡𝜓𝑛,𝑘1,2(𝑡)=𝜓𝑛,𝑘2,2(𝑡).(A.3)

B. Formulae of the Basis for the Doubly-Integrated Wiener Process

For the doubly-integrated Wiener process, the construction of the three-dimensional process involves a family of three 3-dimensional functions, which constitutes the columns of a 3×3 matrix that we denote by 𝜓. This basis has again a simple expression when 𝑚𝑛,𝑘 is the middle of the interval [𝑙𝑛,𝑘,𝑟𝑛,𝑘]:𝜓𝑛,𝑘1,1(𝑡)=𝜓𝑛,𝑘1,2(𝑡)560𝑟𝑛,𝑘𝑙𝑛,𝑘5/2𝑡𝑙𝑛,𝑘3𝑙2𝑛,𝑘7𝑙𝑛,𝑘𝑡+5𝑙𝑛,𝑘𝑟𝑛,𝑘25𝑡𝑟𝑛,𝑘+16𝑡2+10𝑟2𝑛,𝑘,560𝑟𝑛,𝑘𝑙𝑛,𝑘5/2𝑟𝑛,𝑘𝑡3𝑟2𝑛,𝑘7𝑟𝑛,𝑘𝑡+5𝑙𝑛,𝑘𝑟𝑛,𝑘25𝑡𝑙𝑛,𝑘+16𝑡2+10𝑙2𝑛,𝑘,𝜓𝑛,𝑘1,2(𝑡)=312𝑟𝑛,𝑘𝑙𝑛,𝑘5/2𝑙𝑛,𝑘𝑡32𝑟𝑛,𝑘+𝑙𝑛,𝑘3𝑡𝑟𝑛,𝑘+𝑙𝑛,𝑘2𝑡,312𝑟𝑛,𝑘𝑙𝑛,𝑘5/2𝑟𝑛,𝑘𝑡3𝑟𝑛,𝑘+2𝑙𝑛,𝑘3𝑡𝑟𝑛,𝑘+𝑙𝑛,𝑘2𝑡,𝜓𝑛,𝑘1,3(𝑡)=16𝑟𝑛,𝑘𝑙𝑛,𝑘5/2𝑡𝑙𝑛,𝑘32𝑡𝑟𝑛,𝑘𝑙𝑛,𝑘𝑟𝑛,𝑘+𝑙𝑛,𝑘3𝑡2𝑟𝑛,𝑘+𝑙𝑛,𝑘2𝑡,16𝑟𝑛,𝑘𝑙𝑛,𝑘5/2𝑟𝑛,𝑘𝑡32𝑡𝑟𝑛,𝑘𝑙𝑛,𝑘𝑟𝑛,𝑘+𝑙𝑛,𝑘3𝑡2𝑟𝑛,𝑘+𝑙𝑛,𝑘2𝑡,𝜓𝑛,𝑘2,1(𝑡)=16𝑟𝑛,𝑘𝑙𝑛,𝑘5/2𝑡𝑙𝑛,𝑘23𝑟𝑛,𝑘+𝑙𝑛,𝑘4𝑡𝑟𝑛,𝑘+𝑙𝑛,𝑘2𝑡,16𝑟𝑛,𝑘𝑙𝑛,𝑘5/2𝑟𝑛,𝑘𝑡23𝑙𝑛,𝑘+𝑟𝑛,𝑘4𝑡𝑟𝑛,𝑘+𝑙𝑛,𝑘2𝑡,𝜓𝑛,𝑘2,2(𝑡)=36𝑟𝑛,𝑘𝑙𝑛,𝑘5/2𝑡𝑙𝑛,𝑘23𝑟𝑛,𝑘+2𝑙𝑛,𝑘5𝑡𝑟𝑛,𝑘+2𝑙𝑛,𝑘3𝑡,36𝑟𝑛,𝑘𝑙𝑛,𝑘5/2𝑟𝑛,𝑘𝑡23𝑙𝑛,𝑘+2𝑟𝑛,𝑘5𝑡𝑙𝑛,𝑘+2𝑟𝑛,𝑘3𝑡,𝜓𝑛,𝑘2,3(𝑡)=16𝑟𝑛,𝑘𝑙𝑛,𝑘5/2𝑡𝑙𝑛,𝑘32𝑡𝑟𝑛,𝑘𝑙𝑛,𝑘𝑟𝑛,𝑘+𝑙𝑛,𝑘3𝑡2𝑟𝑛,𝑘+𝑙𝑛,𝑘2𝑡,16𝑟𝑛,𝑘𝑙𝑛,𝑘5/2𝑟𝑛,𝑘𝑡32𝑡𝑟𝑛,𝑘𝑙𝑛,𝑘𝑟𝑛,𝑘+𝑙𝑛,𝑘3𝑡2𝑟𝑛,𝑘+𝑙𝑛,𝑘2𝑡,𝜓𝑛,𝑘3,1(𝑡)=53𝑟𝑛,𝑘𝑙𝑛,𝑘5/2𝑡𝑙𝑛,𝑘4𝑙2𝑛,𝑘17𝑙𝑛,𝑘𝑡+9𝑙𝑛,𝑘𝑟𝑛,𝑘15𝑡𝑟𝑛,𝑘+3𝑟2𝑛,𝑘+16𝑡2,53𝑟𝑛,𝑘𝑙𝑛,𝑘5/2𝑟𝑛,𝑘𝑡4𝑟2𝑛,𝑘17𝑡𝑟𝑛,𝑘+9𝑙𝑛,𝑘𝑟𝑛,𝑘15𝑡𝑙𝑛,𝑘+3𝑙2𝑛,𝑘+16𝑡2,𝜓𝑛,𝑘3,2(𝑡)=3(𝑟𝑙)5/2(𝑙𝑡)(𝑟+𝑙2𝑡)(𝑟+4𝑙5𝑡),3(𝑟𝑙)5/2(𝑟𝑡)(𝑟+𝑙2𝑡)(𝑙+4𝑟5𝑡),𝜓𝑛,𝑘3,3(𝑡)=13(𝑟𝑙)5/2(𝑙𝑡)19𝑙256𝑙𝑡+18𝑙𝑟24𝑡𝑟+3𝑟2+40𝑡2,13(𝑟𝑙)5/2(𝑟𝑡)19𝑟256𝑟𝑡+18𝑙𝑟24𝑡𝑙+3𝑙2+40𝑡2.(B.1) Notice again that the basis functions satisfy the relationships𝑑𝑑𝑡𝜓𝑛,𝑘𝑖,𝑗(𝑡)=𝜓𝑛,𝑘𝑖+1,𝑗(𝑡)(B.2) for 𝑖{1,2} and 𝑗{1,2,3}. These functions also form a triorthogonal basis of functions, which makes it easy to simulate sample paths of the doubly-integrated Wiener process, as show in Figure 5.

C. Properties of the Lift Operators

This appendix is devoted to the proofs of the properties of the lift operator enumerated in Proposition 6.3. The proposition is split into three lemmas for the sake of clarity.

Lemma C.1. The operator 𝛼,𝛽𝐺 is a linear measurable bijection. Moreover, for every 𝑁>0, the function 𝛼,𝛽𝐺𝑁=𝑃𝑁𝛼,𝛽𝐺𝐼𝑁𝜉Ω𝑁𝜉Ω𝑁 is a finite-dimensional linear operator, whose matrix representation is triangular in the natural basis of 𝜉Ω𝑁 and whose eigenvalues 𝛼,𝛽𝜈𝑛,𝑘 are given by 𝛼,𝛽𝜈𝑛,𝑘=𝑔𝛼𝑚𝑛,𝑘𝑔𝛽𝑚𝑛,𝑘𝛽𝑀𝑛,𝑘𝛼𝑀𝑛,𝑘,0𝑛𝑁,0𝑘<2𝑁1.(C.1)
Eventually, 𝛼,𝛽𝐺 is abounded operator for the spectral norm with 𝛼,𝛽𝐺2=sup𝑛sup𝑘𝛼,𝛽𝜈𝑛,𝑘sup𝑔𝛼inf𝑔𝛽sup𝑓𝛽inf𝑓sup𝑓2𝛼inf𝑓2𝛽<,(C.2) and the determinant of 𝛼,𝛽𝐺𝑁 denoted by 𝛼,𝛽𝐽𝑁 admits a limit when 𝑁 tends to infinity: lim𝑁det𝛼,𝛽𝐺𝑁=lim𝑁𝛼,𝛽𝐽𝑁=exp1210(𝛼(𝑡)𝛽(𝑡))𝑑𝑡=𝛼,𝛽𝐽.(C.3)

Proof. All these properties are deduced from the properties of the functions Δ and Ψ derived previously.(i)𝛼,𝛽𝐺=𝛽Δ𝛼Ψ is a linear measurable bijection of 𝜉Ω due to the composed application of two linear bijective measurable functions 𝛼Δ𝑥Ω𝜉Ω and 𝛼Ψ𝜉Ω𝑥Ω.(ii)Since we have the expressions of the matrices of the finite-dimensional linear transformations, it is easy to write the linear transformation of 𝛼,𝛽𝐺𝑁 on the natural basis 𝑒𝑛,𝑘 as 𝛼,𝛽𝐺𝑁(𝜉)𝑛,𝑘=𝑈𝛽𝛿𝑛,𝑘(𝑡)𝛼Ψ(𝜉)(𝑡)𝑑𝑡=(𝑝,𝑞)𝑁𝑈𝛽𝛿𝑛,𝑘(𝑡)𝛼𝜓𝑝,𝑞(𝑡)𝑑𝑡𝜉𝑝,𝑞,(C.4) leading to the coefficient expression 𝛼,𝛽𝐺𝑛,𝑘𝑝,𝑞=𝑈𝛽𝛿𝑛,𝑘(𝑡)𝛼𝜓𝑝,𝑞(𝑡)𝑑𝑡=𝑖,𝑗𝑁𝛽Δ𝑛,𝑘𝑖,𝑗𝛼Ψ𝑖,𝑗𝑝,𝑞,(C.5) where we have dropped the index 𝑁 since the expression of the coefficients does not depend on it. We deduce from the form of the matrices 𝛽Δ𝑁 and 𝛼Ψ𝑁 that the application 𝛼,𝛽𝐺𝑁 has a matrix representation in the basis 𝑒𝑛,𝑘 of the form 𝛼,𝛽𝐺𝑁=𝛼,𝛽𝐺0,00,0𝛼,𝛽𝐺0,01,0𝛼,𝛽𝐺1,01,0𝛼,𝛽𝐺0,02,0𝛼,𝛽𝐺1,02,0𝛼,𝛽𝐺2,02,0𝛼,𝛽𝐺0,02,1𝛼,𝛽𝐺1,02,1𝛼,𝛽𝐺2,12,1𝛼,𝛽𝐺0,03,0𝛼,𝛽𝐺1,03,0𝛼,𝛽𝐺2,03,0𝛼,𝛽𝐺3,03,0𝛼,𝛽𝐺0,03,1𝛼,𝛽𝐺1,03,1𝛼,𝛽𝐺2,03,1𝛼,𝛽𝐺3,13,1𝛼,𝛽𝐺0,03,2𝛼,𝛽𝐺1,03,2𝛼,𝛽𝐺2,13,2𝛼,𝛽𝐺3,23,2𝛼,𝛽𝐺0,03,3𝛼,𝛽𝐺1,03,3𝛼,𝛽𝐺2,13,3𝛼,𝛽𝐺3,33,3,(C.6) where we only represent the nonzero terms.The eigenvalues of the operator are therefore the diagonal elements 𝛼,𝛽𝐺𝑛,𝑘𝑛,𝑘 that are easily computed from the expressions of the general term of the matrix: 𝛼,𝛽𝐺𝑛,𝑘𝑛,𝑘=𝛽Δ𝑛,𝑘𝑛,𝑘𝛼𝜓𝑛,𝑘𝑛,𝑘=𝛽𝑀𝑛,𝑘𝑔𝛽𝑚𝑛,𝑘𝑔𝛼𝑚𝑛,𝑘𝛼𝑀𝑛,𝑘=𝑔𝛼𝑚𝑛,𝑘𝑔𝛽𝑚𝑛,𝑘𝛽𝑀𝑛,𝑘𝛼𝑀𝑛,𝑘.(C.7)(iii)From the expression of 𝛼𝑀𝑛,𝑘=(𝛼(𝑟)𝛼(𝑚))(𝛼(𝑚)𝛼(𝑙))/(𝛼(𝑟)𝛼(𝑙)), we deduce the inequalities sup𝑓inf𝑓2𝛼2𝑛+1𝛼𝑀𝑛,𝑘sup𝑓inf𝑓2𝛼2𝑛+1,(C.8) from which follows the given upper-bound to the singular values.(iv)𝛼,𝛽𝐺𝑁 is a finite-dimensional triangular linear matrix in the basis {𝑒𝑛,𝑘}. Its determinant is simply given as the product 𝛼,𝛽𝐽𝑁=𝑁𝑛=00𝑘<2𝑛1𝛼,𝛽𝐺𝑛,𝑘𝑛,𝑘=𝑁𝑛=00𝑘<2𝑛1𝛼,𝛽𝜈𝑛,𝑘,(C.9) where we noticed that the eigenvalues 𝛼,𝛽𝜈𝑛,𝑘 are of the form 𝛼,𝛽𝜈𝑛,𝑘=𝑔𝛼𝑚𝑛,𝑘𝑔𝛽𝑚𝑛,𝑘𝛽𝑀𝑛,𝑘𝛼𝑀𝑛,𝑘.(C.10) Since, for every 𝑁>0, we have 𝛼,𝛽𝐺𝑁=𝜁,𝛽𝐺𝑁𝛼,𝜁𝐺𝑁, which entails that 𝛼,𝛽𝐽𝑁=𝜁,𝛽𝐽𝑁(𝜁,𝛼𝐽𝑁)1, it is enough to show that we have lim𝑁𝛼,0𝐽𝑁=exp1210𝛼(𝑡)𝑑𝑡.(C.11) Now, writing for every 0𝑠<𝑡1 the quantity 𝛼𝒱𝑡,𝑠=𝑡𝑠Γ(𝑢)𝑒2𝑡𝑢𝛼(𝑣)𝑑𝑣𝑑𝑢,(C.12) we have 𝑔𝛼𝑚𝑛,𝑘𝛼𝑀𝑛,𝑘2=𝛼𝒱𝑙𝑛,𝑘,𝑚𝑛,𝑘𝛼𝒱𝑚𝑛,𝑘,𝑟𝑛,𝑘𝛼𝒱𝑙𝑛,𝑘,𝑟𝑛,𝑘,(C.13) so that 𝛼,0𝐽𝑁 is a telescoping product that can be written as 𝛼,0𝐽𝑁2=2𝑁𝑘=0𝛼𝒱𝑘2𝑁,(𝑘+1)2𝑁0𝒱𝑘2𝑁,(𝑘+1)2𝑁.(C.14) If 𝛼 is Hölder continuous, there exist 𝛿>0 and 𝐶>0 such that sup0𝑠,𝑡1||𝛼(𝑡)𝛼(𝑠)|||𝑡𝑠|𝛿<𝐶,(C.15) and introducing, for any 0𝑠<𝑡1, the quantity 𝒬𝑡,𝑠, 𝒬𝑡,𝑠=𝑒𝑡𝑠𝛼(𝑣)𝑑𝑣𝛼𝒱𝑡,𝑠0𝒱𝑡,𝑠=|||||𝑡𝑠Γ(𝑢)𝑒𝑡𝑢𝛼(𝑣)𝑑𝑣𝑢𝑠𝛼(𝑣)𝑑𝑣𝑑𝑢𝑡𝑠Γ(𝑢)𝑑𝑢|||||,(C.16) we have that 𝒬𝑡,𝑠𝒬𝑡,𝑠𝒬𝑡,𝑠 with 𝒬𝑡,𝑠=𝑡𝑠Γ(𝑢)𝑒(𝐶/(1+𝛿))((𝑡𝑢)1+𝛿+(𝑢𝑠)1+𝛿)𝑑𝑢𝑡𝑠Γ(𝑢)𝑑𝑢,𝒬𝑡,𝑠=𝑡𝑠Γ(𝑢)𝑒(𝐶/(1+𝛿))((𝑡𝑢)1+𝛿+(𝑢𝑠)1+𝛿)𝑑𝑢𝑡𝑠Γ(𝑢)𝑑𝑢.(C.17) After expanding the exponential in the preceding definitions, we have 𝒬𝑡,𝑠12𝐶sup0𝑡1Γ(𝑡)inf0𝑡1Γ(𝑡)(𝑡𝑠)(1+𝛿)(1+𝛿)(2+𝛿)+𝑜(𝑡𝑠)(1+𝛿),𝒬𝑡,𝑠1+2𝐶sup0𝑡1Γ(𝑡)inf0𝑡1Γ(𝑡)(𝑡𝑠)(1+𝛿)(1+𝛿)(2+𝛿)+𝑜(𝑡𝑠)(1+𝛿),(C.18) now, from 2𝑁𝑘=0𝒬𝑘2𝑁,(𝑘+1)2𝑁=1+𝑜2𝑁,(C.19) we can directly conclude that lim𝑁𝛼,0𝐽𝑁=𝑒(1/2)10𝛼(𝑡)𝑑𝑡lim𝑁2𝑁𝑘=0𝒬𝑘2𝑁,(𝑘+1)2𝑁=𝑒(1/2)10𝛼(𝑡)𝑑𝑡.(C.20)

Notice that, if 𝛼=𝛽, 𝛼,𝛼𝐺 is the identity and 𝛼,𝛼𝐽=1 as expected.

Similar properties are now proved for the process lift operator 𝛼,𝛽𝐻.

Lemma C.2. The function 𝛼,𝛽𝐻 is a linear measurable bijection.
Moreover, for every 𝑁>0, the function 𝛼,𝛽𝐻𝑁=𝑃𝑁𝛼,𝛽𝐻𝐼𝑁𝑥Ω𝑁𝑥Ω𝑁 is a finite-dimensional linear operator, whose matrix representation is triangular in the natural basis of 𝑥Ω𝑁 and whose eigenvalues are given by 𝛽,𝛼𝜈𝑛,𝑘=𝛼,𝛽𝜈𝑛,𝑘1=𝑔𝛽𝑚𝑛,𝑘𝑔𝛼𝑚𝑛,𝑘𝛼𝑀𝑛,𝑘𝛽𝑀𝑛,𝑘.(C.21) Eventually, 𝛼,𝛽𝐻 is a bounded operator for the spectral norm with 𝛼,𝛽𝐻2=𝛽,𝛼𝐺2=sup𝑛sup𝑘𝛽,𝛼𝜈𝑛,𝑘sup𝑔𝛽inf𝑔𝛼sup𝑓inf𝑓𝛽sup𝑓2𝛽inf𝑓2𝛼<,(C.22) and the determinant of 𝛼,𝛽𝐻𝑁 admits a limit when 𝑁 tends to infinity: lim𝑁det𝛼,𝛽𝐻=exp1210(𝛽(𝑡)𝛼(𝑡))𝑑𝑡=𝛽,𝛼𝐽.(C.23)

Proof. (i)The function 𝛼,𝛽𝐻=𝛽Ψ𝛼Δ is a linear measurable bijection of 𝑥Ω onto 𝑥Ω because 𝛼Δ𝑥Ω𝜉Ω and 𝛼Ψ𝜉Ω𝑥Ω are linear bijective measurable functions.(ii)We write the linear transformation of 𝛼,𝛽𝐻 for 𝑥 in 𝑥Ω as 𝛼,𝛽𝐻[𝑥](𝑡)=(𝑛,𝑘)𝛽𝜓𝑛,𝑘(𝑡)𝑈𝛼𝛿𝑛,𝑘(𝑠)𝑥(𝑠)𝑑𝑠=𝑈(𝑛,𝑘)𝛽𝜓𝑛,𝑘(𝑡)𝛼𝛿𝑛,𝑘(𝑡)𝑥(𝑡)𝑑𝑡.(C.24) If we denote the class of 𝑥 in 𝑥Ω by 𝑥={𝑥𝑖,𝑗}={𝑥(𝑚𝑖,𝑗)}, (𝑖,𝑗)𝑁, we can write (C.24) as 𝛼,𝛽𝐻𝑁(𝑥)𝑖,𝑗=(𝑘,𝑙)𝑁(𝑝,𝑞)𝑁𝛽Ψ𝑖,𝑗𝑝,𝑞𝛼Δ𝑝,𝑞𝑘,𝑙𝑥𝑘,𝑙,(C.25) from which we deduce the expression of the coefficients of the matrix 𝛼,𝛽𝐻𝑁: 𝛼,𝛽𝐻𝑖,𝑗𝑘,𝑙=(𝑝,𝑞)𝑁𝛽Ψ𝑖,𝑗𝑝,𝑞𝛼Δ𝑝,𝑞𝑘,𝑙,(C.26) where as usual we drop the index 𝑁. Because of the the form of the matrices 𝛼Δ𝑁 and 𝛽Ψ𝑁, the matrix 𝛼,𝛽𝐻𝑁 in the basis 𝑓𝑖,𝑗 has the following triangular form: 𝛼,𝛽𝐻𝑁=𝛼,𝛽𝐻0,00,0𝛼,𝛽𝐻0,01,0𝛼,𝛽𝐻1,01,0𝛼,𝛽𝐻0,02,0𝛼,𝛽𝐻1,02,0𝛼,𝛽𝐻2,02,0𝛼,𝛽𝐻0,02,1𝛼,𝛽𝐻1,02,1𝛼,𝛽𝐻2,12,1𝛼,𝛽𝐻0,03,0𝛼,𝛽𝐻1,03,0𝛼,𝛽𝐻2,03,0𝛼,𝛽𝐻3,03,0𝛼,𝛽𝐻0,03,1𝛼,𝛽𝐻1,03,1𝛼,𝛽𝐻2,03,1𝛼,𝛽𝐻3,13,1𝛼,𝛽𝐻0,03,2𝛼,𝛽𝐻1,03,2𝛼,𝛽𝐻2,13,2𝛼,𝛽𝐻3,23,2𝛼,𝛽𝐻0,03,3𝛼,𝛽𝐻1,03,3𝛼,𝛽𝐻2,13,3𝛼,𝛽𝐻3,33,3.(C.27) From the matrix representations 𝛼Δ and 𝛽Ψ, the diagonal terms of 𝛼,𝛽𝐻 read 𝛼,𝛽𝐻𝑖,𝑗𝑖,𝑗=𝛽Ψ𝑖,𝑗𝑖,𝑗𝛼Δ𝑖,𝑗𝑖,𝑗=𝛽𝜓𝑖,𝑗𝑚𝑖,𝑗𝛼𝑀𝑖,𝑗𝑔𝛼𝑚𝑖,𝑗=𝑔𝛽𝑚𝑖,𝑗𝑔𝛼𝑚𝑖,𝑗𝛼𝑀𝑖,𝑗𝛽𝑀𝑖,𝑗=𝛽,𝛼𝜈𝑖,𝑗=𝛼,𝛽𝜈𝑖,𝑗1.(C.28)(iii)The upper bound directly follows from the fact that 𝛽,𝛼𝜈𝑖,𝑗=(𝛼,𝛽𝜈𝑖,𝑗)1.(iv) Since 𝛽,𝛼𝜈𝑖,𝑗=(𝛼,𝛽𝜈𝑖,𝑗)1, the value of the determinant of 𝛼,𝛽𝐻𝑁 is clearly the inverse of the determinant of 𝛼,𝛽𝐺𝑁, so that lim𝑁det(𝛼,𝛽𝐻𝑁)=(𝛼,𝛽𝐽)1=𝛽,𝛼𝐽.

Note that Lemma C.2 directly follows from the fact that 𝛼Ψ and 𝛼Δ are inverse of each other and admit a triangular matrix representation. More precisely, when restricted to the finite-dimensional case, we have set the following properties.

Properties C.3. We have the following set of properties in terms of matrix operations(i)𝛼,𝛽𝐻𝑁=𝛽Ψ𝑁𝛼Δ𝑁 and 𝛼,𝛽𝐺𝑁=𝛽Δ𝑁𝛼Ψ𝑁,(ii)𝛼,𝛽𝐻𝑁=𝜁,𝛽𝐻𝑁𝛼,𝜁𝐻𝑁 and 𝛼,𝛽𝐺𝑁=𝜁,𝛽𝐺𝑁𝛼,𝜁𝐺𝑁,(iii)𝛼,𝛽𝐻1𝑁=𝛽,𝛼𝐻𝑁 and 𝛼,𝛽𝐺1𝑁=𝛽,𝛼𝐺𝑁,(iv)𝛼,𝛽𝐻𝑁𝛼Ψ𝑁=𝛽Ψ𝑁 and 𝛼,𝛽𝐺𝑁𝛼Δ𝑁=𝛽Δ𝑁,(v)𝛼,𝛽𝐻𝑁𝑇𝛽Δ𝑁𝑇=𝛼Δ𝑁𝑇 and 𝛼,𝛽𝐺𝑁𝑇𝛽Φ𝑁𝑇=𝛼Φ𝑁𝑇.

Proof. (i)Let us write 𝜉𝑉𝑁=𝑁𝑛=0vect({𝑒𝑛,𝑘}0𝑘<2𝑛1) and 𝑥𝑉𝑁=𝑁𝑛=0vect({𝑓𝑖,𝑗}0𝑗<2𝑖1). Since 𝛼Ψ and 𝛽Ψ project the flag 𝜉𝑉0𝜉𝑉1𝜉𝑉𝑁 onto the flag 𝑥𝑉0𝑥𝑉1𝑥𝑉𝑁, and since conversely 𝛼Δ and 𝛼Δ project the flag 𝑥𝑉0𝑥𝑉1𝑥𝑉𝑁 onto the flag 𝜉𝑉0𝜉𝑉1𝜉𝑉𝑁, we can write𝛼,𝛽𝐻𝑁=𝑥𝑃𝑁𝛽Ψ𝛼Δ𝑥𝐼𝑁=𝑥𝑃𝑁𝛽Ψ𝑥𝐼𝑁𝑥𝑃𝑁𝛼Δ𝑥𝐼𝑁=𝛽Ψ𝑁𝛼Δ𝑁,𝛼,𝛽𝐺𝑁=𝜉𝑃𝑁𝛽Δ𝛼Ψ𝜉𝐼𝑁=𝜉𝑃𝑁𝛽Δ𝜉𝐼𝑁𝜉𝑃𝑁𝛼Ψ𝜉𝐼𝑁=𝛽Δ𝑁𝛼Ψ𝑁.(C.29)(ii)We have𝜁,𝛽𝐻𝑁𝛼,𝜁𝐻𝑁=𝛽Ψ𝑁𝜁Δ𝑁𝜁Ψ𝑁𝛼Δ𝑁=𝛽Ψ𝑁𝐼𝑑𝜉Ω𝑁𝛼Δ𝑁=𝛼,𝛽𝐻𝑁,𝜁,𝛽𝐺𝑁𝛼,𝜁𝐺𝑁=𝛽Δ𝑁𝜁Ψ𝑁𝜁Δ𝑁𝛼Ψ𝑁=𝛽Δ𝑁𝐼𝑑𝑥Ω𝑁𝛼Ψ𝑁.=𝛼,𝛽𝐻𝑁.(C.30)(iii)We have𝛼,𝛽𝐻𝑁𝛽,𝛼𝐻𝑁=𝛼,𝛼𝐻𝑁=𝐼𝑑𝜉Ω𝑁,𝛼,𝛽𝐺𝑁𝛽,𝛼𝐺𝑁=𝛼,𝛼𝐺𝑁=𝐼𝑑𝜉Ω𝑁.(C.31)(iv)We have𝛼,𝛽𝐻𝑁𝛼Ψ𝑁=𝛽Ψ𝑁𝛼Δ𝑁𝛼Ψ𝑁=𝛽Ψ𝑁𝐼𝑑𝜉Ω𝑁=𝛽Ψ𝑁,𝛼,𝛽𝐺𝑁𝛼Δ𝑁=𝛽Δ𝑁𝛼Ψ𝑁𝛼Δ𝑁=𝛽Δ𝑁𝐼𝑑𝑥Ω𝑁=𝛽Δ𝑁.(C.32)(v)We have𝛼,𝛽𝐻𝑁𝑇𝛽Δ𝑁𝑇=𝛼Δ𝑁𝑇𝛽Ψ𝑁𝑇𝛽Δ𝑁𝑇=𝛼Δ𝑁𝑇𝐼𝑑𝑥Ω𝑁=𝛼Δ𝑁𝑇,𝛼,𝛽𝐺𝑁𝑇𝛽Ψ𝑁𝑇=𝛼Ψ𝑁𝑇𝛽Δ𝑁𝑇𝛽Ψ𝑁𝑇=𝛼Ψ𝑁𝑇𝐼𝑑𝜉Ω𝑁=𝛼Ψ𝑁𝑇.(C.33)

D. Construction and Coefficient Applications

In this appendix, we provide the proofs of the main properties used in the paper regarding the construction and the coefficient applications.

D.1. The Construction Application

We start by addressing the case of the construction application introduced in Section 3.2.1.

We start by proving Proposition 3.11.

Proof. For the sake of simplicity, we will denotes for any function 𝐀[0,1]𝑚×𝑑, the uniform norm as |𝐀|=sup0𝑡1|𝐀(𝑡)|, where |𝐀(𝑡)|=sup0𝑖<𝑚(𝑑10|𝐴𝑖,𝑗(𝑡)|) is the operator norm induced by the uniform norms. We will also denote the 𝑖th line of 𝐀 by 𝑙𝑖(𝐀) (it is a 𝑑-valued function) and the 𝑗th column of 𝐀 by 𝑐𝑗(𝐀).
Let 𝝃𝜉Ω be fixed. These coefficients induce a sequence of continuous functions 𝝍𝑁(𝝃) through the action of the sequence of the partial construction applications. To prove that this sequence converges towards a continuous function, we show that it uniformly converges, which implies the result of the proposition using the fact that a uniform limit of continuous functions is a continuous function. Moreover, since the functions take values in 𝑑, which is a complete space, we show that for any sequence of coefficients 𝜉𝜉Ω, the sequence of functions 𝜓𝑁(𝑡) constitutes a Cauchy sequence for the uniform norm.
By definition of 𝜉Ω, for every 𝝃 in 𝜉Ω, there exist 𝛿<1 and 𝑛𝜉 such that, for every 𝑛>𝑛𝜉, we have sup0𝑘<2𝑛1||𝝃𝑛,𝑘||<2𝑛𝛿/2,(D.1) which implies that for, 𝑁>𝑛𝜉, we have ||𝚿𝑁(𝝃)(𝑡)𝚿𝑛𝜉(𝝃)(𝑡)||(𝑛,𝑘)𝑁𝑛𝜉||𝝍𝑛,𝑘(𝑡)𝝃𝑛,𝑘||𝑛=𝑛𝜉2𝑛𝛿/2||𝝍𝑛,𝑘||.(D.2)
We therefore need to upperbound the uniform norm of the function 𝝍𝑛,𝑘. To this purpose, we use the definition of 𝝍𝑛,𝑘 given by (3.28): 𝝍𝑛,𝑘(𝑡)=𝑔(𝑡)𝑡0𝑓(𝑠)𝚽𝑛,𝑘(𝑠)𝑑𝑠.(D.3) The coefficient in position (𝑖,𝑗) of the integral term in the right-hand side of the previous inequality can be written as a function of the lines and columns of 𝑓 and Φ𝑛,𝑘 and can be upperbounded using the Cauchy-Schwarz inequality on 𝐿2([0,1],𝑑) as follows: 𝟙[0,𝑡]𝑐𝑖𝐟𝑇,𝑐𝑗𝜙𝑛,𝑘=𝑈𝟙[0,𝑡]𝑆n,𝑘(𝑠)𝑙𝑖(𝐟)(𝑠)𝑐𝑗𝜙𝑛,𝑘(𝑠)𝑑𝑠𝟙[0,𝑡]𝑆𝑛,𝑘𝑙𝑖(𝐟)2𝑐𝑗(𝜙𝑛,𝑘)2.(D.4) Since the columns of Φ𝑛,𝑘 form an orthogonal basis of functions for the standard scalar product in 𝐿2([0,1],𝑑) (see Proposition 3.5), we have 𝑐𝑗(𝜙𝑛,𝑘)2=1. Moreover, since 𝐟 is bounded continuous on [0,1], we can define constants 𝐾𝑖=sup0𝑡1𝑙𝑖(𝐟)(𝑡)< and write 𝟙[0,𝑡]𝑆𝑛,𝑘𝑙𝑖(𝐟)2=𝑈𝟙[0,𝑡]𝑆𝑛,𝑘(𝑠)𝑙𝑖(𝐟)(𝑠)𝑇𝑙𝑖(𝐟)(𝑠)𝑑𝑠𝑈𝟙[0,𝑡]𝑆𝑛,𝑘(𝑠)𝐾2𝑖𝑑𝑠=2𝑛+1𝐾2𝑖<.(D.5) Setting 𝐾=max0𝑖<𝑑𝐾𝑖, for all (𝑛,𝑘) in , the 𝑑×𝑑-valued functions 𝜿𝑛,𝑘(𝑡)=𝑈𝟙[0,𝑡](𝑠)𝐟(𝑠)𝜙𝑛,𝑘(𝑠)𝑑𝑠(D.6) satisfy 𝜿𝑛,𝑘=sup0𝑡1|𝜿𝑛,𝑘(𝑡)|𝐾2(𝑛+1)/2.
Moreover, since 𝐠 is also bounded continuous on [0,1], there exists 𝐿 such that 𝐠=sup0𝑡1|𝐠(𝑡)|𝐿, and we finally have, for all 0𝑡1, 𝝍𝑛,𝑘𝐠𝜿𝑛,𝑘𝐿𝐾2(𝑛+1)/2.(D.7)
Now using this bound and (D.2), we have ||𝚿𝑁(𝝃)(𝑡)𝚿𝑛𝜉(𝝃)(𝑡)||(𝑛,𝑘)𝑁𝑛𝜉||𝝍𝑛,𝑘(𝑡)𝝃𝑛,𝑘||𝐿𝐾2𝑛=𝑛𝜉2(𝛿1)/2𝑛,(D.8) and since 𝛿<1, for the continuous functions 𝑡Ψ𝑁𝑡(𝝃) forms a uniformly convergent sequence of functions for the 𝑑-dimensional uniform norm. This sequence therefore converges towards a continuous function, and Ψ is well defined on 𝜉Ω and takes values in 𝐶0([0,1],𝑑).

This proposition being proved, we dispose of the map Ψ=lim𝑁Ψ𝑁. We now turn to prove different useful properties on this function. We denote by (𝐶0([0,1],𝑑)) the Borelian sets of the 𝑑-dimensional Wiener space 𝐶0([0,1],𝑑).

Lemma D.2. The function Ψ(𝜉Ω,(𝜉Ω))(𝐶0([0,1],𝑑),(𝐶0([0,1],𝑑))) is a linear injection.

Proof. The application Ψ is clearly linear. The injective property simply results from the existence of the dual family of distributions 𝜹𝑛,𝑘. Indeed, for every 𝝃,𝝃 in 𝜉Ω, we have that Ψ(𝜉)=Ψ(𝜉) entails, that for all 𝑛,𝑘,𝝃𝑛,𝑘=𝒫(𝜹𝑛,𝑘,Ψ(𝜉))=𝒫(𝜹𝑛,𝑘,Ψ(𝜉))=𝝃𝑛,𝑘.

In the one-dimensional case, as mentioned in the main text, because the uniform convergence of the sample paths is preserved as long as 𝛼 is continuous and Γ is nonzero through (D.8), the definition 𝑥Ω does not depend on 𝛼 or Γ and the space 𝑥Ω is large enough to contain reasonably regular functions.

Proposition D.3. In the one-dimensional case, the space 𝑥Ω contains the space of uniformly Hölder continuous functions 𝐻 defined as 𝐻=𝑥𝐶[0,1]𝛿>0,sup0𝑠,𝑡1||𝑥(𝑡)𝑥(𝑠)|||𝑡𝑠|𝛿<+.(D.9)

Remark D.3. This point can be seen as a direct consequence of the characterization of the local Hölder exponent of a continuous real function in terms of the asymptotic behavior of its coefficients in the decomposition on the Schauder basis [63].

Proof. To underline that we place ourselves in the one-dimensional case, we drop the bold notations that indicate multidimensional quantities. Supposing that 𝑥 is uniformly Hölder continuous for a given 𝛿>0, there always exists 𝜉 such that Ψ𝑁(𝜉) coincides with 𝑥 on 𝐷𝑁: it is enough to take 𝜉 such that, for all (𝑛,𝑘) in 𝑁, 𝜉𝑛,𝑘=(𝛿𝑛,𝑘,𝑥). We can further write for 𝑛>0𝑥,𝛿𝑛,𝑘=𝑀𝑛,𝑘𝑥𝑚𝑛,𝑘𝑔𝑚𝑛,𝑘𝐿𝑛,𝑘𝑥𝑙𝑛,𝑘𝑔𝛼𝑙𝑛,𝑘+𝑅𝑛,𝑘𝑥𝑟𝑛,𝑘𝑔𝑟𝑛,𝑘,=𝐿𝑛,𝑘𝑥𝑚𝑛,𝑘𝑔𝑚𝑛,𝑘𝑥𝑙𝑛,𝑘𝑔𝑙𝑛,𝑘+𝑅𝑛,𝑘𝑥𝑚𝑛,𝑘𝑔𝑚𝑛,𝑘𝑥𝑟𝑛,𝑘𝑔𝑟𝑛,𝑘.(D.10) For a given function 𝛼, posing 𝑁𝛼=sup0𝑡1𝑓𝛼(𝑡)/inf0𝑡1𝑓2𝛼(𝑡), we have 𝛼𝑀𝑛,𝑘𝑁𝛼2(𝑛+1)/2,𝛼𝐿𝑛,𝑘𝑁𝛼2(𝑛1)/2,𝛼𝑅𝑛,𝑘𝑁𝛼2(𝑛1)/2.(D.11) Moreover, if 𝛼 is in 𝐻, it is straightforward to see that 𝑔𝛼 has a continuous derivative. Then, since 𝑥 is 𝛿-Hölder, for any 𝜖0, there exists 𝐶>0 such that |𝑡𝑠|𝜖 entails that ||||𝑥(𝑡)𝑔(𝑡)𝑥(𝑠)𝑔(𝑠)||||𝐶𝜖𝛿,(D.12) from which we directly deduce ||𝜉𝑛,𝑘||𝑁𝛼𝐶22𝑛((1/2)2𝛿).(D.13) This demonstrates that {𝜉𝑛,𝑘} belongs to 𝜉Ω and ends the proof.

We equip the space 𝑥Ω with the topology induced by the uniform norm on 𝐶0([0,1],𝑑). As usual, we denote by (𝑥Ω) the corresponding Borelian sets. We now show Proposition 3.12.

Proposition D.4. The function Ψ(𝜉Ω,(𝜉Ω))(𝑥Ω,(𝑥Ω)) is a bounded continuous bijection.

Proof. Consider an open ball 𝑥𝐵(𝑥,𝜖) of 𝑥Ω of radius 𝜖. If we take 𝑀=𝐿𝐾/2 as defined in (D.8), we can choose a real 𝛿>0 such that 𝛿<𝜖𝑀𝑛=02𝑛/21.(D.14) Let us consider 𝝃 in 𝜉Ω such that Ψ(𝜉)=𝑥. Then, by (D.8), we immediately have that, for all 𝝃 in the ball of radius 𝜉𝐵(𝜉,𝛿) of 𝜉Ω,Ψ(𝝃𝝃)𝜖. This shows that Ψ1(𝑥𝐵(𝑥,𝜖)) is open and that Ψ is continuous for the 𝑑-dimensional uniform norm topology.

D.2. The Coefficient Application

In this section of the appendix, we show some useful properties of the coefficient application introduced in Section 3.2.2.

Lemma D.5. The function Δ(𝐶0([0,1],𝑑),(𝐶0([0,1],𝑑)))(𝜉Ω,(𝜉Ω)) is a measurable linear injection.

Proof. (i)The function Δ is clearly linear.(ii)To prove that Δ is injective, we show that for 𝐱 and 𝐲 in 𝐶0([0,1],𝑑), 𝐱𝐲 implies that Δ(𝐱)Δ(𝑦). To this end, we fix 𝐱𝐲 in 𝐶0([0,1],𝑑) equipped with the uniform norm and consider the continuous function 𝐝𝑁(𝑡)=(𝑛,𝑘)𝑁𝝍𝑛,𝑘(𝑡)𝚫(𝐱)𝑛,𝑘𝚫(𝐲)𝑛,𝑘.(D.15) This function coincides with 𝐱𝐲 on every dyadic number in 𝐷𝑁 and has zero value if Δ(𝐱)=Δ(𝐲). Since 𝐱𝐲, there exists 𝑠 in]0,1[such that 𝐱(𝑠)𝐲(𝑠), and by continuity of 𝐱𝐲, there exists an 𝜀>0 such that 𝐱𝐲 on the ball]𝑠𝜀,𝑠+𝜀[. But, for 𝑁 large enough, there exists 𝑘, 0k<2𝑁1 such that |𝑠𝑘2𝑁|<𝜀. We then necessarily have that Δ(𝑓)Δ(𝑔); otherwise, we would have 𝑑𝑁(𝑘2𝑁)=(𝐱𝐲)(𝑘2𝑁)=0, which would contradict the choice of 𝜀.(iii)Before proving the measurability of Δ, we need the following observation. Consider for 𝑁>0, the finite-dimensional linear function Δ𝑁𝐶0[0,1],𝑑𝑑2𝑁1,𝐱𝚫𝑁(𝑥)=𝚫(𝑥)𝑁,𝑘(𝑁,𝑘)𝑁.(D.16) Since for all (𝑁,𝑘), the matrices 𝐌𝑁,𝑘, 𝐑𝑁,𝑘, 𝐋𝑁,𝑘 are all bounded, the function Δ𝑁(𝐶0([0,1],𝑑),(𝐶0([0,1],𝑑)))((𝑑)2𝑁1,((𝑑)2𝑁1)) is a continuous linear application. To show that the function Δ is measurable, it is enough to show that the pre-image by Δ of the generative cylinder sets of (𝜉Ω) belongs to (𝐶0([0,1],𝑑)).For any 𝑁0, take an arbitrary Borel set 𝐵=(𝑛,𝑘)𝑁𝐵𝑛,𝑘𝑑𝑁(D.17) and define the cylinder set 𝒞𝑁(𝐵) as 𝒞𝑁(𝐵)=𝝃𝜉Ω(𝑛,𝑘)𝐼𝑁,𝝃𝑛,𝑘𝐵𝑛,𝑘,(D.18) and we write the collection of cylinder sets 𝐶 as 𝐶=𝑛0𝐶𝑁with𝐶𝑁=𝐵𝑑𝑁𝒞𝑁(𝐵).(D.19) We proceed by induction of 𝑁 to show that the preimage by Δ of any cylinder set in 𝐶 is in (𝐶0([0,1],𝑑)). For 𝑁=0, a cylinder set of 𝐶0 is of the form 𝐵0,0 in (𝑑), Δ1(𝐵)={𝐱𝐶0([0,1],𝑑)𝐱(1)𝐋𝑇0,0𝐠1(𝑟0,0)(𝐵0,0)}, which is measurable for being a cylinder set of (𝐶0([0,1],𝑑)). Suppose now that, for 𝑁>0, for any set 𝐴 in 𝐶𝑁1, the set Δ1(𝐴) is measurable. Then, considering a set 𝐴 in 𝐶𝑁, there exists 𝐵 in ((𝑑)𝑁) such that 𝐴=𝒞𝑁(𝐵). Define 𝐴 in 𝐶𝑁 such that 𝐴=𝒞𝑁1(𝐵), where 𝐵=(𝑛,𝑘)𝑁1𝐵𝑛,𝑘,(D.20) and remark that 𝐴=𝒞𝑁(𝐵)𝐴=𝒞𝑁(𝐵). Clearly, we have that 𝐴=𝐴𝐷, where we have defined the cylinder set 𝐷 as 𝐷=𝒞𝐼𝑁(𝑁,𝑘)𝑁,𝑘𝐵𝑁,𝑘.(D.21) Having defined the function Δ𝑁, we now have Δ1(𝐴)=Δ1(𝐴𝐷)=Δ1(𝐴)Δ1(𝐷)=Δ1(𝐴)Δ1𝑁(𝐷). Because of the continuity of Δ𝑁, Δ1𝑁(𝐷) is a Borel set of (𝐶0([0,1],𝑑)). Since, by hypothesis of recurrence, Δ1(𝐴) is in (𝐶0([0,1],𝑑)), Δ1(𝐴) is also in (𝐶0([0,1],𝑑)) as the intersection of two Borel sets. The proof of the measurability of Δ is complete.

We now demonstrate Theorem 3.14.

Proposition D.6. The function Δ(𝑥Ω,(𝑥Ω))(𝜉Ω,(𝜉Ω)) is a measurable linear bijection whose inverse is Ψ=Δ1.

Proof. Let 𝑥𝑥Ω be a continuous function. We have 𝚿(𝚫(𝑥))(𝑡)=(𝑛,𝑘)𝝍𝑛,𝑘(𝑡)𝚫𝑛,𝑘=(𝑛,𝑘)𝝍𝑛,𝑘(𝑡)𝒫𝜹𝑛,𝑘,𝑥.(D.22) This function is equal to 𝑥(𝑡) for any 𝑡𝐷, the set of dyadic numbers. Since 𝐷 is dense in [0,1] and both 𝑥 and Ψ(Δ(𝑥)) are continuous, the two functions, coinciding on the dyadic numbers, are equal for the uniform distance, and hence Ψ(Δ(𝑥))=𝑥.

E. Itô Formula

In this section, we provide rigorous proofs of Proposition 6.1 and Theorem 6.2 related to the Itô formula.

Proposition E.1 (integration by parts). Let (𝑋𝑡) and (𝑌𝑡) be two one-dimensional Gauss-Markov processes starting from zero. Then one has the following equality in law: 𝑋𝑡𝑌𝑡=𝑡0𝑋𝑠𝑑𝑌𝑠+𝑡0𝑌𝑠𝑑𝑋𝑠,(E.1) where 𝑡0𝐴𝑠𝑑𝐵𝑠 two stochastic processes denotes for 𝐴𝑡 and 𝐵𝑡 the Stratonovich integral. In terms of the Itô integral, this formula is written as 𝑋𝑡𝑌𝑡=𝑡0𝑋𝑠𝑑𝑌𝑠+𝑡0𝑌𝑠𝑑𝑋𝑠+𝑋,𝑌𝑡,(E.2) where the brackets denote the mean quadratic variation.

Proof. We assume that 𝑋 and 𝑌 satisfy the equations: 𝑑𝑋𝑡=𝛼𝑋(𝑡)𝑋𝑡+Γ𝑋(𝑡)𝑑𝑊𝑡,𝑑𝑌𝑡=𝛼𝑌(𝑡)𝑋𝑡+Γ𝑌(𝑡)𝑑𝑊𝑡,(E.3) and we introduce the functions 𝑓𝑋, 𝑓𝑌, 𝑔𝑋, and 𝑔𝑌 such that 𝑋𝑡=𝑔𝑋(𝑡)𝑡0𝑓𝑋(𝑠)and 𝑌𝑡=𝑔𝑌(𝑡)𝑡0𝑓𝑌(𝑠).
We define (𝑋𝜓𝑛,𝑘)(𝑛,𝑘) and (𝑌𝜓𝑛,𝑘)(𝑛,𝑘), the construction bases of the processes 𝑋 and 𝑌. Therefore, using Theorem 4.5, there exist (𝑋Ξ𝑛,𝑘)(𝑛,𝑘) and (𝑌Ξ𝑝,𝑞)(𝑝,𝑞) standard normal independent variables such that 𝑋=(𝑛,𝑘)𝑋𝜓𝑛,𝑘𝑋Ξ𝑛,𝑘 and 𝑌=(𝑝,𝑞)𝑌𝜓𝑛,𝑘𝑌Ξ𝑛,𝑘 and we know that the processes 𝑋 and 𝑌 are almost-surely uniform limits when 𝑁 of the processes 𝑋𝑁 and 𝑌𝑁 defined as the partial sums: 𝑋𝑁=(𝑛,𝑘)𝑁𝑋𝜓𝑛,𝑘𝑋Ξ𝑛,𝑘,𝑌𝑁=(𝑝,𝑞)𝑁𝑌𝜓𝑛,𝑘𝑌Ξ𝑛,𝑘.(E.4) Using the fact that the functions 𝑋𝜓𝑛,𝑘 and 𝑌𝜓𝑛,𝑘 have piecewise continuous derivatives, we have 𝑋𝑁𝑡𝑌𝑁𝑡=(𝑛,𝑘)𝐼𝑁(𝑝,𝑞)𝐼𝑁𝑋𝜓𝑛,𝑘(𝑡)𝑌𝜓𝑝,𝑞(𝑡)𝑋Ξ𝑛,𝑘𝑌Ξ𝑝,𝑞=(𝑛,𝑘)𝐼𝑁(𝑝,𝑞)𝐼𝑁𝑋Ξ𝑛,𝑘𝑌Ξ𝑝,𝑞𝑡0𝑑𝑑𝑠𝑋𝜓𝑛,𝑘(𝑠)𝑌𝜓𝑝,𝑞(𝑠)𝑑𝑡=(𝑛,𝑘)𝐼𝑁(𝑝,𝑞)𝐼𝑁𝑋Ξ𝑛,𝑘𝑌Ξ𝑝,𝑞𝑡0𝑋𝜓𝑛,𝑘(𝑠)𝑌𝜓𝑝,𝑞(𝑠)+𝑋𝜓𝑛,𝑘(𝑠)𝑌𝜓𝑝,𝑞(𝑠)𝑑𝑠.(E.5) Therefore, we need to evaluate the piecewise derivative of the functions 𝑋𝜓𝑛,𝑘 and 𝑌𝜓𝑛,𝑘. We know that 1𝑓𝑋𝑋𝜓𝑛,𝑘𝑔𝑋(𝑡)=𝑋𝜙𝑛,𝑘(𝑡),(E.6) which entails that 𝑋𝜓𝑛,𝑘=𝛼𝑋𝑋𝜓𝑛,𝑘+𝑔𝑋𝑓𝑋𝑋𝜙𝑛,𝑘=𝛼𝑋𝑋𝜓𝑛,𝑘+Γ𝑋𝑋𝜙𝑛,𝑘(E.7) and similarly so for the process 𝑌. Therefore, we have 𝑋𝑁𝑌𝑁=𝐴𝑛,𝑘𝑝,𝑞+𝐵𝑛,𝑘𝑝,𝑞+𝐶𝑛,𝑘𝑝,𝑞+𝐷𝑛,𝑘𝑝,𝑞,(E.8) with 𝐴𝑡=(𝑛,𝑘)𝐼𝑁(𝑝,𝑞)𝐼𝑁𝑡0𝛼𝑋(𝑠)𝑋𝜓𝑛,𝑘(𝑠)𝑌𝜓𝑝,𝑞(𝑠)𝑑𝑠𝑋Ξ𝑛,𝑘𝑌Ξ𝑝,𝑞,𝐵𝑡=(𝑛,𝑘)𝐼𝑁(𝑝,𝑞)𝐼𝑁𝑡0Γ𝑋(𝑠)𝑋𝜙𝑛,𝑘(𝑠)𝑌𝜓𝑝,𝑞(𝑠)𝑑𝑠𝑋Ξ𝑛,𝑘𝑌Ξ𝑝,𝑞,𝐶𝑡=(𝑛,𝑘)𝐼𝑁(𝑝,𝑞)𝐼𝑁𝑡0𝛼𝑌(𝑠)𝑋𝜓𝑛,𝑘(𝑠)𝑌𝜓𝑝,𝑞(𝑠)𝑑𝑠𝑋Ξ𝑛,𝑘𝑌Ξ𝑝,𝑞,𝐷𝑡=(𝑛,𝑘)𝐼𝑁(𝑝,𝑞)𝐼𝑁𝑡0Γ𝑌(𝑠)𝑌𝜙𝑛,𝑘(𝑠)𝑋𝜓𝑝,𝑞(𝑠)𝑑𝑠𝑋Ξ𝑛,𝑘𝑌Ξ𝑝,𝑞.(E.9) We easily compute 𝐴𝑡+𝐶𝑡=𝑡0𝛼𝑋(𝑠)+𝛼𝑌(𝑠)𝑋𝑁(𝑠)𝑌𝑁(𝑠)𝑑𝑠.(E.10) For 𝑡[0,1], as it is our case, 𝑋𝑁(𝑠) and 𝑌𝑁(𝑠) are both almost surely finite for all 𝑡 in [0,1]. For almost all 𝑌𝜉 and 𝑌𝜉 drawn with respect to the law of the Gaussian infinite vector Ξ, we therefore have, by the Lebesgue dominated convergence theorem, that this integral converges almost surely towards 𝑡0𝛼𝑋(𝑠)+𝛼𝑌(𝑠)𝑋(𝑠)𝑌(𝑠)𝑑𝑠.(E.11) The other two terms 𝐵𝑡 and 𝐷𝑡 necessitate a more thorough analysis, and we treat them as follows. Let us start by considering the first one of this term: 𝐵𝑡=10𝟙[0,𝑡](𝑠)Γ𝑋(𝑠)(𝑛,𝑘)𝐼𝑁(𝑝,𝑞)𝐼𝑁𝑋𝜙𝑛,𝑘(𝑠)𝑌𝜓𝑝,𝑞(s)𝑋Ξ𝑛,𝑘𝑌Ξ𝑝,𝑞𝑑𝑠=𝑡𝑖𝒟𝑁{1}𝑡𝑖+1𝑡𝑖𝟙[0,𝑡](𝑠)Γ𝑋(𝑠)(𝑛,𝑘)𝑁𝑋𝜙𝑛,𝑘(𝑠)𝑋Ξ𝑛,𝑘𝑌𝑁(𝑠)𝑑𝑠=𝑡𝑖𝒟𝑁{1}𝑡𝑖+1𝑡𝑖𝟙[0,𝑡](𝑠)Γ𝑋(𝑠)1𝑓𝑋(𝑠)𝑋𝑁𝑔𝑋(𝑠)𝑌𝑁(𝑠)𝑑𝑠.(E.12) Let us now have a closer look at the process 𝑋𝑁𝑡 for 𝑡[𝑡𝑖,𝑡𝑖+1] where [𝑡𝑖,𝑡𝑖+1]=𝑆𝑁,𝑖 for 𝑖 such that (𝑁,𝑖). Because of the structure of our construction, we have 𝑌𝑁(𝑡)=𝑔𝑌(𝑡)𝑔𝑌𝑡𝑖𝑌𝑡𝑖+1𝑌(𝑡)𝑌𝑡𝑖+1𝑌𝑡𝑖𝑌𝑡𝑖+𝑔𝑌(𝑡)𝑔𝑌𝑡𝑖+1𝑌(𝑡)𝑌𝑡𝑖𝑌𝑡𝑖+1𝑌𝑡𝑖𝑌𝑡𝑖+1,(E.13)1𝑓𝑋(𝑡)𝑋𝑁𝑔𝑋(𝑡)=𝑓𝑋(𝑡)𝑋𝑡𝑖+1𝑋𝑡𝑖𝑋𝑡𝑖+1𝑔𝑋𝑡𝑖+1𝑋𝑡𝑖𝑔𝑋𝑡𝑖.(E.14) We therefore have 𝑡𝑖+1𝑡𝑖1[0,𝑡](𝑠)Γ𝑋(𝑠)𝑓𝑋(𝑠)𝑑𝑑𝑠𝑋𝑁(𝑠)𝑔𝑋(𝑠)𝑌𝑁(𝑠)𝑑𝑠=𝑡𝑖+1𝑡𝑖1[0,𝑡](𝑠)Γ𝑋(𝑠)𝑓𝑋(𝑠)𝑋𝑡𝑖+1𝑋𝑡𝑖𝑔𝑌(𝑠)×𝑌𝑡𝑖+1𝑌(𝑠)𝑌𝑡𝑖+1𝑌𝑡𝑖𝑌𝑡𝑖𝑔𝑌𝑡𝑖+𝑌(𝑠)𝑌𝑡𝑖𝑌𝑡𝑖+1𝑌𝑡𝑖𝑌t𝑖+1𝑔𝑌𝑡𝑖+1𝑑𝑠𝑋𝑡𝑖+1𝑔𝑋𝑡𝑖+1𝑋𝑡𝑖𝑔𝑋𝑡𝑖=𝑣𝑖(𝑡)𝑁𝑌𝑁𝑡𝑖+𝑤𝑖(𝑡)𝑁𝑌𝑁𝑡𝑖+1𝑋𝑡𝑖+1𝑔𝑋𝑡𝑖+1𝑋𝑡𝑖𝑔𝑋𝑡𝑖(E.15) with 𝑣𝑁𝑖(𝑡)=𝑡𝑖+1𝑡𝑖𝟙[0,𝑡](𝑠)Γ𝑋(𝑠)𝑓𝑋(𝑠)𝑋𝑡𝑖+1𝑋𝑡𝑖𝑔𝑌(𝑠)g𝑌𝑡𝑖𝑌𝑡𝑖+1𝑌(𝑠)𝑌𝑡𝑖+1𝑌𝑡𝑖𝑑𝑠,𝑤𝑁𝑖(𝑡)=𝑡𝑖+1𝑡𝑖𝟙[0,𝑡](𝑠)Γ𝑋(𝑠)𝑓𝑋(𝑠)𝑋𝑡𝑖+1𝑋𝑡𝑖𝑔𝑌(𝑠)𝑔𝑌𝑡𝑖+1𝑌(𝑠)𝑌𝑡𝑖𝑌𝑡𝑖+1𝑌𝑡𝑖𝑑𝑠.(E.16) Let us denote by 𝛿𝑁 the time step of the partition 𝛿𝑁=max𝑡𝑖𝒟𝑁{1}(𝑡𝑖+1𝑡𝑖), which is smaller than 𝜌𝑁 with 𝜌(0,1) from the assumption made in Section 3.2.1. Moreover, we know that the functions 𝑔𝑋=𝑔𝑌, 𝑋 and 𝑌 are continuously differentiable, and since Γ𝑋 and Γ𝑌 are 𝛿-Hölder, so are 𝑓𝑋 and 𝑓𝑌. When 𝑁 (i.e., when 𝛿𝑁0), using the Taylor and Hölder expansions for the differential functions, we can further evaluate the integrals we are considering. Let us first assume that 𝑡>𝑡𝑖+1. We have 𝑣𝑖(𝑡)𝑁=𝑡𝑖+1𝑡𝑖Γ𝑋(𝑠)𝑓𝑋(𝑠)𝑋𝑡𝑖+1𝑋𝑡𝑖𝑔𝑌(𝑠)𝑔𝑌𝑡𝑖𝑌𝑡𝑖+1𝑌(𝑠)𝑌𝑡𝑖+1𝑌𝑡𝑖𝑑𝑠=Γ𝑋𝑡𝑖1+𝑂𝛿𝛿𝑁𝑓𝑋𝑡𝑖1+𝑂𝛿𝛿𝑁𝑓𝑋𝑡𝑖2𝑡𝑖+1𝑡𝑖1+𝑂𝛿𝑁×𝑔𝑌𝑡𝑖1+𝑂𝛿𝑁𝑔𝑌𝑡𝑖𝑡𝑖+1𝑡𝑖𝑓𝑌𝑡𝑖2𝑡i+1𝑠1+𝑂𝛿𝑁𝑓𝑌𝑡𝑖2𝑡𝑖+1𝑡𝑖1+𝑂𝛿𝑁𝑑𝑠=Γ𝑋𝑡𝑖𝑓𝑋𝑡𝑖𝑡𝑖+1𝑡𝑖𝑡𝑖+1𝑡𝑖𝑡𝑖+1𝑠𝑡𝑖+1𝑡𝑖𝑑𝑠1+𝑂𝛿𝑁+𝑂𝛿𝛿𝑁=12𝑔𝑋𝑡𝑖+𝑂𝛿𝑁+𝛿𝛿𝑁.(E.17) Similarly, we show that 𝑤𝑁𝑖(𝑡)=(1/2)𝑔𝑋(𝑡𝑖)+𝑂(𝛿𝑁+𝛿𝛿𝑁) when 𝑁. If 𝑡<𝑡𝑖, we have 𝑣𝑁𝑖(𝑡)=𝑤𝑁𝑖(𝑡)=0 and for 𝑡 in [𝑡𝑖0,𝑡𝑖0+1) we have 𝑣𝑁𝑖0(𝑡)=𝑔𝑋𝑡𝑖02𝑡𝑖0+1𝑡𝑡𝑖0+1𝑡𝑖02+𝑂𝛿𝑁+𝛿𝛿𝑁,𝑤𝑁𝑖0(𝑡)=𝑔𝑋𝑡𝑖02𝑡𝑡𝑖0𝑡𝑖0+1𝑡𝑖02+𝑂𝛿𝑁+𝛿𝛿𝑁=𝑣𝑁𝑖0(𝑡)+𝑂𝛿𝑁+𝛿𝛿𝑁.(E.18) We then finally have 𝐵𝑡=𝑡𝑖𝒟𝑁;𝑡𝑖+1𝑡𝑔𝑋𝑡𝑖2𝑌𝑁𝑡𝑖+𝑌𝑁𝑡𝑖+1𝑋𝑁𝑡𝑖+1𝑔𝑋𝑡𝑖+1𝑋𝑁𝑡𝑖𝑔𝑋𝑡𝑖+𝑔𝑋𝑡𝑖02𝑡𝑖0+1𝑡𝑡𝑖0+1𝑡𝑖02𝑌𝑁𝑡𝑖0+𝑌𝑁𝑡𝑋𝑁𝑡𝑖+1𝑔𝑋(𝑡)𝑋𝑁𝑡𝑖0𝑔𝑋𝑡𝑖0+𝑂𝛿𝑁+𝛿𝛿𝑁.(E.19) Moreover, we observe that the process 𝑋𝑡/𝑔𝑋(𝑡)=𝑡𝑖+1𝑡𝑖𝑓𝑋(𝑠)𝑑𝑊𝑠 is a martingale, and by definition of the Stratonovich integral for martingale processes, we have 𝐵𝑡𝑁𝑡0𝑔𝑋(𝑠)𝑌𝑠𝑑𝑋𝑠𝑔𝑋(𝑠)=𝑡0Γ𝑋(𝑠)𝑌𝑠𝑑𝑊(𝑠),(E.20) where is used to denote the Stratonovich stochastic integral and the limit is taken in distribution. Notice that the fact that the sum converges towards the Stratonovich integral does not depend on the type of sequence of partition chosen which can be different from the dyadic partition. Putting all these results together, we obtain the equality in law: 𝑋𝑡𝑌𝑡=𝑡0𝛼𝑋(𝑠)𝑋𝑠𝑌𝑠𝑑𝑠+𝑡0Γ𝑋(𝑠)𝑌𝑠𝑑𝑊𝑠+𝑡0𝛼𝑌(𝑠)𝑋𝑠𝑌𝑠𝑑𝑠+𝑡0Γ𝑌(𝑠)𝑋𝑠𝑑𝑊𝑠,(E.21) which is exactly the integration by parts formula we were searching for. The integration by parts formula for the Itô stochastic integral directly comes from the relationship between the Stratonovich and Itô stochastic integrals.

Theorem E.1 (Itô). Let 𝑋 be a Gauss-Markov process and 𝑓 in 𝐶2(). The process 𝑓(𝑋𝑡) is a Markov process and satisfies the relation 𝑓𝑋𝑡=𝑓𝑋0+𝑡0𝑓𝑋𝑠𝑑𝑋𝑠+12𝑡0𝑓𝑋𝑠𝑑𝑋𝑠.(E.22)

Proof. The integration by parts formula directly implies the Itô formula through a density argument as follows. Let 𝒜 be the set of functions 𝑓𝐶2([0,1],) such that (E.22) is true. It is clear that 𝒜 is a vector space. Moreover, because of the result of Proposition 6.1, the space 𝒜 is an algebra. Since all constant functions and the identity function 𝑓(𝑥)=𝑥 trivially belong to 𝒜, the algebra 𝒜 contains all polynomial functions.
Let now 𝑓𝐶2([0,1],). There exists a sequence of polynomials 𝑃𝑘 such that 𝑃𝑘 (resp., 𝑃𝑘,𝑃𝑘) uniformly converges towards 𝑓 (resp., 𝑓,𝑓). Let us denote by 𝑈𝑛 the sequence of stopping times: 𝑈𝑛=inf𝑡[0,1];||𝑋𝑡||>𝑛.(E.23) This sequence grows towards infinity. We have 𝑃𝑘𝑋𝑡𝑈𝑛𝑃𝑘𝑋0=𝑡0𝑃𝑘𝑋𝑠𝟙[0,𝑈𝑛](𝑠)𝑑𝑋𝑠+12𝑡0𝑃𝑘𝑋𝑠𝟙[0,𝑈𝑛](𝑠)𝑑𝑋𝑠.(E.24) On the interval [0,𝑈𝑛], we have 𝑋𝑡𝑛, which allows to use the Lebesgue dominated convergence theorem on each term of the equality. We have 𝔼||||𝑡0𝑃𝑘𝑋𝑠𝟙[0,𝑈𝑛](𝑠)𝑑𝑋𝑠𝑡0𝐹𝑋𝑠𝟙[0,𝑈𝑛](𝑠)𝑑𝑋𝑠||||2=𝔼𝑡0||𝑃𝑘(𝑋𝑠)𝐹𝑋𝑠||2𝟙[0,𝑈𝑛](𝑠)𝑑𝑋𝑠,(E.25) which converges towards zero because of the Lebesgue theorem for the Steljes integration. The same argument directly applies to the other term. Therefore, letting 𝑘, we proved Itô formula for 𝑋𝑡𝑈𝑛, and eventually letting 𝑛, we obtain the desired formula.

F. Trace Class Operator

In this section, we demonstrate Theorem 6.6, which proves the instrumental to extend the finite-dimensional change of variable formula to the infinite-dimensional case. The proof relies on the following lemma.

Lemma F.1. The operator 𝛼,𝛽𝑆𝐼𝑑𝑙2()𝑙2() is isometric to the operator 𝛼,𝛽𝑅𝑙2()𝑙2() defined by 𝛼,𝛽𝑅[𝑥]=10𝛼,𝛽𝑅(𝑡,𝑠)𝑥(𝑠)𝑑𝑠,(F.1) with the kernel 𝛼,𝛽𝑅(𝑡,𝑠)=(𝛼(𝑡𝑠)𝛽(𝑡𝑠))𝑓𝛼(𝑡𝑠)𝑓𝛼(𝑡𝑠)+𝑓𝛼(𝑡)1𝑡𝑠(𝛼(𝑢)𝛽(𝑢))2𝑓𝛼2(𝑢)𝑑𝑢𝑓𝛼(𝑠).(F.2)

Proof. Notice first that (𝛼(𝑡𝑠)𝛽(𝑡𝑠))𝑓𝛼(𝑡𝑠)𝑓𝛼(𝑡𝑠)=𝟏{𝑠<𝑡}(𝛼(𝑡)𝛽(𝑡))𝑓𝛼(𝑠)𝑓𝛼(𝑡)+𝟏{𝑠𝑡}(𝛼(𝑠)𝛽(𝑠))𝑓𝛼(𝑡)𝑓𝛼(𝑠),(F.3) which leads to writing in 𝐿2[0,1], for any (𝑛,𝑘) and (𝑝,𝑞) in 𝛼𝜙𝑛,𝑘,𝛼,𝛽𝑅𝛼𝜙𝑝,𝑞=10𝛼,𝛽𝑅(𝑡,𝑠)𝛼𝜙𝑛,𝑘(𝑡)𝛼𝜙𝑝,𝑞(𝑠)𝑑𝑡𝑑𝑠=𝐴𝑛,𝑘𝑝,𝑞+𝐵𝑛,𝑘𝑝,𝑞+𝐶𝑛,𝑘𝑝,𝑞,(F.4) with 𝐴𝑛,𝑘𝑝,𝑞=10𝛼(𝑡)𝛽(𝑡)𝑓𝛼(𝑡)𝛼𝜙𝑛,𝑘(𝑡)𝑡0𝑓𝛼(𝑠)𝛼𝜙𝑝,𝑞(𝑠)𝑑𝑠𝑑𝑡,=10𝛼(𝑡)𝛽(𝑡)Γ(𝑡)𝛼𝜙𝑛,𝑘(𝑡)𝛼𝜓𝑝,𝑞(𝑡)𝑑𝑡,𝐵𝑛,𝑘𝑝,𝑞=10𝛼𝑓(𝑡)𝜙𝑛,𝑘(𝑡)1𝑡𝛼(𝑠)𝛽(𝑠)𝑓𝛼(𝑠)𝛼𝜙𝑝,𝑞(𝑠)𝑑𝑠𝑑𝑡,=10𝛼(𝑡)𝛽(𝑡)Γ(𝑡)𝛼𝜓𝑛,𝑘(𝑡)𝛼𝜙𝑝,𝑞(𝑡)𝑑𝑡,𝐶𝑛,𝑘𝑝,𝑞=10𝛼𝜙𝑛,𝑘(𝑡)𝑓𝛼(𝑡)𝑠,𝑡(𝛼(𝑢)𝛽(𝑢))2𝑓𝛼2(𝑢)𝑑𝑢𝑓(𝑠)𝛼𝜙𝑝,𝑞(𝑠)𝑑𝑡𝑑𝑠,=10(𝛼(𝑢)𝛽(𝑢))2𝑓𝛼2(𝑢)×𝟏[0,𝑢](𝑡)𝑓𝛼(𝑡)𝛼𝜙𝑝,𝑞(𝑡)𝟏[0,𝑢](𝑠)𝑓𝛼(𝑠)𝛼𝜙𝑝,𝑞(𝑠)𝑑𝑡𝑑𝑠𝑑𝑢,=10(𝛼(𝑢)𝛽(𝑢))2Γ(𝑢)𝛼𝜓𝑛,𝑘(𝑡)𝛼𝜓𝑝,𝑞(𝑡)𝑑𝑡𝑑𝑠.(F.5) This proves that (𝛼𝜙𝑛,𝑘,𝛼,𝛽𝑅[𝛼𝜙𝑝,𝑞])=[𝛼,𝛽𝑆𝐼𝑑]𝑛,𝑘𝑝,𝑞. Therefore, if we denote the isometric linear operator 𝛼Φ𝑙2()𝐿2(),𝜉𝛼Φ[𝜉]=𝑛=00𝑘<2𝑛1𝛼𝜙𝑛,𝑘𝜉𝑛,𝑘,(F.6) we clearly have 𝛼Φ𝑇𝛼,𝛽𝑅𝛼Φ=𝛼,𝛽𝑆𝐼𝑑 with 𝛼Φ𝑇=𝛼Φ1.

We now proceed to demonstrate that 𝛼,𝛽𝑆𝐼𝑑 is a trace class operator.

Proof of Theorem 6.6. Since the kernel 𝛼,𝛽𝑅(𝑡,𝑠) is integrable in 𝐿2([0,1]×[0,1]), the integral operator 𝛼,𝛽𝑅𝐿2[0,1]𝐿2[0,1] is a Hilbert-Schmidt operator and thus is compact. Moreover, it is a trace class operator since we have Tr𝛼,𝛽𝑅=10(𝛼(𝑡)𝛽(𝑡))𝑑𝑡+10𝑓𝛼2(𝑡)1𝑡(𝛼(𝑡)𝛽(𝑡))2𝑓𝛼2(𝑢)𝑑𝑠𝑑𝑡=10(𝛼(𝑡)𝛽(𝑡))𝑑𝑡+10𝛼(𝑡)𝑓𝛼(𝑡)2(𝛼(𝑡)𝛽(𝑡))2𝑑𝑡.(F.7) Since 𝛼,𝛽𝑆𝐼𝑑 and 𝛼,𝛽𝑅 are isometric through 𝛼Φ, the compactness of 𝛼,𝛽𝑆𝐼𝑑 is equivalent to the compactness of 𝛼,𝛽𝑅. Moreover, the traces of both operators coincide: 𝑛=00𝑘<2𝑛1𝛼,𝛽𝑆𝑛,𝑘𝑛,𝑘=𝑛=00𝑘<2𝑛110𝛼𝜙𝑛,𝑘(𝑡)𝛼𝜙𝑛,𝑘(𝑠)𝛼,𝛽𝑅(𝑡,𝑠)𝑑𝑠𝑑𝑡,=10𝑛=00𝑘<2𝑛1𝛼𝜙𝑛,𝑘(𝑡)𝛼𝜙𝑛,𝑘(𝑠)𝛼,𝛽𝑅(𝑡,𝑠)𝑑𝑠𝑑𝑡,=10𝛼,𝛽𝑅(𝑡,𝑡)𝑑𝑠𝑑𝑡,(F.8) using the result of Corollary 3.6.

G. Girsanov Formula

In this section we provide the quite technical proof of Lemma 6.8 which is useful in proving the Girsanov formula.

Lemma G.1. The positive definite quadratic form on 𝑙2()×𝑙2() associated with operator 𝛼,𝛽𝑆𝐼𝑑𝑙2()𝑙2() is well defined on 𝜉Ω. Moreover for all 𝜉Ω, 𝜉,𝛼,𝛽𝑆𝐼𝑑𝜉Ω(𝜉)=210𝛼(𝑡)𝛽(𝑡)𝑓2(𝑡)𝛼𝑋𝑡(𝜉)𝑔𝛼(𝑡)𝑑𝛼𝑋𝑡(𝜉)𝑔𝛼(𝑡)+10(𝛼(𝑡)𝛽(𝑡))2𝑓2(𝑡)𝛼𝑋𝑡(𝜉)𝑔𝛼(𝑡)2𝑑𝑡,(G.1) where 𝛼𝑋𝑡(𝜉)=𝛼Ψ(𝜉) and refers to the Stratonovich integral and the equality is true in law.

Proof. The proof of this lemma uses quite similar materials to those used in the proof of the Itô theorem. However, since this result is central for giving insight on the way our geometric considerations relate to the Girsanov theorem, we provide the detailed proof here.
Consider 𝜉 in 𝜉Ω, denote 𝜉𝑁=𝜉𝑃𝑁(𝜉), and write 𝜉𝑁,𝛼,𝛽𝑆𝑁𝐼𝑑𝜉Ω𝑁𝜉𝑁=(𝑛,𝑘)𝑁(𝑝,𝑞)𝑁𝐴𝑛,𝑘𝑝,𝑞+𝐵𝑛,𝑘𝑝,𝑞𝜉𝑛,𝑘𝜉𝑝,𝑞,(G.2) where we have posited 𝐴𝑛,𝑘𝑝,𝑞=210𝛼(𝑡)𝛽(𝑡)Γ(𝑡)𝛼𝜙𝑛,𝑘(𝑡)𝛼𝜓𝑝,𝑞(𝑡)+𝛼𝜙𝑝,𝑞(𝑡)𝛼𝜓𝑛,𝑘(𝑡)𝑑𝑡,𝐵𝑛,𝑘𝑝,𝑞=10(𝛼(𝑡)𝛽(𝑡))2Γ(𝑡)𝛼𝜓𝑛,𝑘(𝑡)𝛼𝜓𝑝,𝑞(𝑡)𝑑𝑡.(G.3) It is easy to see, using similar arguments to those in the proof of the integration by parts formula, Proposition 6.1: 𝐴𝑁(𝜉)=(𝑛,𝑘)𝑁(𝑝,𝑞)𝑁𝐴𝑛,𝑘𝑝,𝑞𝜉𝑛,𝑘𝜉𝑝,𝑞=210𝛼(𝑡)𝛽(𝑡)Γ(𝑡)𝛼𝑋𝑁𝑡(𝜉)𝑓𝛼(𝑡)𝑑𝑑𝑡𝛼𝑋𝑁𝑡(𝜉)𝑔𝛼(𝑡)𝑑𝑡,(𝑛,𝑘)𝑁(p,𝑞)𝑁𝐵𝑛,𝑘𝑝,𝑞𝜉𝑛,𝑘𝜉𝑝,𝑞=10(𝛼(𝑡)𝛽(𝑡))2Γ(𝑡)𝛼𝑋𝑁𝑡(𝜉)2𝑑𝑡.(G.4) Because of the uniform convergence property of 𝑋𝑁 towards 𝑋 and the fact that it has almost surely bounded sample paths, the latter sum converges towards 10(𝛼(𝑡)𝛽(𝑡))2𝑓2(𝑡)𝛼𝑋𝑡(𝜉)𝑔𝛼(𝑡)2𝑑𝑡.(G.5) Now writing quantity 𝐴𝑁(𝜉) as the sum of elementary integrals between the points of discontinuity 𝑡𝑖=𝑖2𝑁, 0𝑖2𝑁, 𝐴𝑁(𝜉)=22𝑁1𝑖=0𝑡𝑖+1𝑡𝑖𝛼(𝑡)𝛽(𝑡)Γ(𝑡)𝛼𝑋𝑁𝑡(𝜉)𝑓(𝑡)𝑑𝑑𝑡𝛼𝑋𝑁𝑡(𝜉)𝑔𝛼(𝑡)𝑑𝑡(G.6) and using the identities of (E.13) and (E.14), we then have 𝐴𝑁(𝜉)=22𝑁1𝑖=0𝑤𝑁𝑖𝛼𝑋𝑡𝑖(𝜉)𝑔𝛼𝑡𝑖+𝑤𝑁𝑖+1𝛼𝑋𝑡𝑖+1(𝜉)𝑔𝛼𝑡𝑖+1𝛼𝑋𝑡𝑖+1(𝜉)g𝛼𝑡𝑖+1𝛼𝑋𝑡𝑖(𝜉)𝑔𝛼𝑡𝑖,(G.7) where we denote 𝑤𝑁𝑖=𝑡𝑖+1𝑡𝑖(𝛼(𝑡)𝛽(𝑡))𝑡𝑖+1(𝑡)𝑑𝑡𝑡𝑖+1𝑡𝑖2,𝑤𝑁𝑖+1=𝑡𝑖+1𝑡𝑖(𝛼(𝑡)𝛽(𝑡))𝑡𝑖+1(𝑡)𝑑𝑡𝑡𝑖+1𝑡𝑖2.(G.8) Let us define the function 𝑤 in 𝐶[0,1] by 𝑤(𝑡)=𝛼(𝑡)𝛽(𝑡)𝑓2(𝑡).(G.9) If 𝛼 and 𝛽 are uniformly 𝛿-Hölder continuous, so is 𝑤. Therefore, there exist an integer 𝑁>0 and a real 𝑀>0 such that if 𝑁>𝑁, for all 0𝑖<2𝑁, we have ||𝑤𝑁𝑖𝑤𝑡𝑖||=||||||𝑡𝑖+1𝑡𝑖𝑤(𝑡)𝑤𝑡𝑖(𝑑/𝑑𝑡)𝑡𝑖+1(𝑡)2𝑡𝑖+1𝑡𝑖2𝑑𝑡||||||𝑀||||||𝑡𝑖+1𝑡𝑖𝑡𝑡𝑖𝛿(𝑑/𝑑𝑡)𝑡𝑖+1(𝑡)2𝑡𝑖+1𝑡𝑖2𝑑𝑡||||||𝑀𝑡𝑡𝑖𝛿+12(𝛿+1)+𝑀|||||𝑡𝑖+1𝑡𝑖𝑡𝑡𝑖𝛿+1𝛿+1𝑡𝑖+1(𝑡)2𝑡𝑖+1𝑡𝑖2𝑑𝑡|||||𝑀𝑡𝑡𝑖𝛿+12(𝛿+1)+𝑀𝑡𝑡𝑖𝛿+22(𝛿+1)(𝛿+2),(G.10) which shows that |𝑤𝑁𝑖𝑤(𝑡𝑖)|=𝑂(2𝑁(1+𝛿)), and similarly |𝑤𝑁𝑖+1𝑤(𝑡𝑖+1)|=𝑂(2𝑁(1+𝛿)) as well. As a consequence, expression Lemma G.1 converges when 𝑁 tends to infinity toward the desired Stratonovich integral.

Acknowledgments

The authors wish to thank Professor Marcelo Magnasco for many illuminating discussions. This work was partially supported by NSF Grant EF-0928723 and ERC Grant NERVI-227747.