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

Discrete stopping times in the lattice of continuous functions

Achintya Raya Polavarapu polavara@ualberta.ca Department of Mathematical and Statistical Sciences, University of Alberta, Edmonton, AB, T6G 2G1, Canada.
Abstract.

A functional calculus for an order complete vector lattice \mathcal{E} was developed by Grobler in [Gro14b] using the Daniell integral. We show that if one represents the universal completion of \mathcal{E} as C(K)C^{\infty}(K), then the Daniell functional calculus for continuous functions is exactly the pointwise composition of functions in C(K)C^{\infty}(K). This representation allows an easy deduction of the various properties of the functional calculus. Afterwards, we study discrete stopping times and stopped processes in C(K)C^{\infty}(K). We obtain a representation that is analogous to what is expected in probability theory.

Key words and phrases:
vector lattice, stochastic process, stopping time, stopped process, universal completion
2010 Mathematics Subject Classification:
46A40, 46E05, 60G20, 60G40.

1. Introduction and Preliminaries

In the first half of this article, we prove some results regarding the functional calculus in vector lattices. Specifically, the main result shows that when represented in C(K)C^{\infty}(K), the Daniell functional calculus introduced by Grobler in [Gro14b] is the pointwise composition of functions for continuous functions. In the second half, we investigate discrete stopping times and their representation in C(K)C^{\infty}(K). In Section 3, we show that discrete stopping times in vector lattices correspond to a certain class of elements in the sup-completion. Moreover, when elements of the vector lattice are viewed as functions in C(K)C^{\infty}(K), the discrete stopping times and stopped processes have a representation similar to what is expected in classical probability. This allows an easy deduction of various properties of discrete stopping times in vector lattices. Furthermore, we provide a generalization of Début theorem to the abstract setting. For more information about stopping times and hitting times in the classical probability theory, we refer the reader to [Kal17].

1.1. C(K)C^{\infty}(K) representation

We refer the reader to [AB03, AB06] for the theory of vector lattices. Given a Hausdorff topological space KK, we will write C(K)C(K) for the set of all continuous functions from KK to \mathbb{R}. The space C(K)C(K) is a vector lattice under pointwise operations, and if KK is an extremally disconnected Hausdorff topological space, then C(K)C(K) is order (or Dedekind) complete. A subset AA of a Hausdorff topological space is nowhere dense if (A¯)=(\overline{A})^{\circ}=\emptyset. This also implies that A\partial A is a nowhere dense set for every closed set AA. A set AA is meagre if it is a union of a sequence of nowhere dense sets, and conversely, a set is co-meagre if its complement is meagre. KK is a Baire space if countable unions of closed sets with empty interior also have empty interior. Every compact Hausdorff space is a Baire space.

Given that KK is an extremally disconnected compact Hausdorff topological space, we write C(K)C^{\infty}(K) for the vector lattice of all continuous functions from KK to [,][-\infty,\infty] that are finite almost everywhere (a.e.), that is, except on a nowhere dense set. Two functions ff and gg in C(K)C^{\infty}(K) are equal, provided that the set of points where their values differ is a nowhere dense set. Scalar multiplication, addition and lattice operations on C(K)C^{\infty}(K) are defined pointwise a.e. It should be noted that C(K)C^{\infty}(K) is an f-algebra with product defined pointwise a.e. Maeda-Ogasawara Theorem states that every Archimedean vector lattice \mathcal{E} may be represented as an order dense sublattice of C(K)C^{\infty}(K), where KK is an extremally disconnected compact Hausdorff topological space. Throughout the paper, we fix an order complete vector lattice \mathcal{E} with a weak unit EE and a Maeda-Ogasawara representation of \mathcal{E} as an order dense ideal in C(K)C^{\infty}(K), where KK is the Stone space of \mathcal{E}, with EE corresponding to 𝟙\mathbbm{1}.

We shall also use the concept of the sup-completion of \mathcal{E} as introduced by [Don82] which we will denote by s\mathcal{E}^{s}. We recall [[PT], Theorem 1] that states that the sup-completion of an order complete vector lattice \mathcal{E} is s={fC(K,¯):fg for some g}\mathcal{E}^{s}=\{f\in C(K,\overline{\mathbb{R}}):f\geq g\text{ for some }g\in\mathcal{E}\}. We refer the reader to [PT] for more properties regarding the representation of s\mathcal{E}^{s}. Every element of the sup-completion can be decomposed into finite and infinite parts [AN22, PT]. In terms of the above representation, this decomposition is as follows: the finite part xx of uu is defined as x=PUux=P_{U}u where UU is the closure of {u<}\{u<\infty\}. The infinite part ww of uu is defined as PUcuP_{U^{c}}u and is the function that vanishes on UU and takes the value \infty on UcU^{c}. Corollary 7 in [Azo19, PT] also gives us that +u+s\mathcal{E}_{+}^{u}\subseteq\mathcal{E}_{+}^{s}.

1.2. Conditional expectation operator

The motivation for the following definitions and theory can be found in [KLW04a, KLW05]. A conditional expectation 𝔽\mathbb{F} defined on \mathcal{E} is an order continuous strictly positive linear projection whose range (𝔽)\mathcal{R}(\mathbb{F}) is an order complete vector lattice of \mathcal{E} with 𝔽E=E\mathbb{F}E=E. Conditional expectation operators satisfy an averaging property [[KLW05], Theorem 5.3] in the sense that if f(𝔽)f\in\mathcal{R}(\mathbb{F}) and gg\in\mathcal{E} with fgfg\in\mathcal{E} then 𝔽(fg)=f𝔽(g)\mathbb{F}(fg)=f\mathbb{F}(g), where the product of two elements is done in C(K)C^{\infty}(K). We note that the range of a conditional expectation operator is a regular sublattice of \mathcal{E}.

Let J+J\subset\mathbb{R}_{+}. A filtration on \mathcal{E} is a family (𝔽t)tJ(\mathbb{F}_{t})_{t\in J} of conditional expectations satisfying 𝔽s=𝔽s𝔽t=𝔽t𝔽s\mathbb{F}_{s}=\mathbb{F}_{s}\mathbb{F}_{t}=\mathbb{F}_{t}\mathbb{F}_{s} for all sts\leq t. We denote the range of 𝔽t\mathbb{F}_{t} by t\mathcal{F}_{t}. A stochastic process in \mathcal{E} is a family (Xt)tJ(X_{t})_{t\in J} where XtX_{t}\in\mathcal{E}, with J+J\subset\mathbb{R}^{+} an interval. The stochastic process (Xt)tJ(X_{t})_{t\in J} is right continuous if o-limstXs=Xt\text{o-}\lim_{s\downarrow t}X_{s}=X_{t}. The stochastic process (Xt)tJ(X_{t})_{t\in J} is adapted to the filtration if XttX_{t}\in\mathcal{F}_{t} for all tJt\in J. If (Xt)(X_{t}) is a stochastic process adapted to (𝔽t)(\mathbb{F}_{t}), we call (Xt,𝔽t)(X_{t},\mathbb{F}_{t}) a super-martingale (respectively sub-martingale) if 𝔽t(Xs)Xt\mathbb{F}_{t}(X_{s})\leq X_{t} (respectively 𝔽t(Xs)Xt\mathbb{F}_{t}(X_{s})\geq X_{t}) for all tst\leq s. The process is called a martingale if it is a sub-martingale and super-martingale.

2. Functional Calculus

The functional calculus developed by Grobler plays a major role in the theory of measure-free probability in vector lattices [AT17, AR18, Gro14a, AN]. In the following subsection, we briefly outline the construction of the functional calculus using the Daniell integral, as shown in [Gro14b].

2.1. Daniell Integral

Denote by F()F(\mathbb{R}) the algebra consisting of all finite unions of disjoint left open right closed intervals (a,b],(a,)(a,b],(a,\infty) and (,b](\infty,b] with a,ba,b\in\mathbb{R}. Let 𝕃\mathbb{L} be the vector lattice of real valued functions of the form:

(1) f=i=1nai𝟙Si,SiF()f=\sum_{i=1}^{n}a_{i}\mathbbm{1}_{S_{i}},S_{i}\in F(\mathbb{R})

where (Si)i=1n(S_{i})_{i=1}^{n} is a partition of \mathbb{R}. The order relation of 𝕃\mathbb{L} is defined by fgf\leq g if f(t)g(t)f(t)\leq g(t) for every tt\in\mathbb{R}.

Definition 1.

A positive linear function I:𝕃I:\mathbb{L}\to\mathcal{E} is called an \mathcal{E}-valued Daniell integral on 𝕃\mathbb{L} whenever, for every sequence (fn)(f_{n}) in 𝕃\mathbb{L} that satisfies fn(t)0f_{n}(t)\downarrow 0 for every tt\in\mathbb{R}, it follows that I(fn)0I(f_{n})\downarrow 0.

We note that II need not be an order continuous operator. Consider I:𝕃I:\mathbb{L}\to\mathbb{R} where I(f)=f(0)I(f)=f(0). Then II is a \mathbb{R}-valued Daniell integral and consider the sequence fn=𝟙(1n,0]f_{n}=\mathbbm{1}_{(-\frac{1}{n},0]}. Then fn0f_{n}\downarrow 0 in 𝕃\mathbb{L} but I(fn)↛0I(f_{n})\not\to 0.

We shall detail the construction of a specific Daniell integral developed by Grobler that shall be useful in defining a functional calculus on \mathcal{E}. For YY\in\mathcal{E} we denote PYP_{Y} to be the band projection associated with the band generated by YY. Fix XX\in\mathcal{E}, then the right continuous spectral system of XX is the increasing right-continuous stochastic process A=(At)tA=(A_{t})_{t\in\mathbb{R}} where At=EP(XtE)+EA_{t}=E-P_{(X-tE)^{+}}E. We denote by AA_{\infty} and AA_{-\infty} respectively, the supremum and infimum of the process. To define the \mathcal{E}-valued Daniell integral, we define a vector lattice measure μA\mu_{A} with respect to (At)t(A_{t})_{t\in\mathbb{R}} as follows:

  • μA(a,b]=AbAa\mu_{A}(a,b]=A_{b}-A_{a} where (a,b]F()(a,b]\in F(\mathbb{R}).

  • For any finite disjoint union of half open intervals, we have μA(i=1nIi)=i=1nμA(Ii)\mu_{A}(\cup_{i=1}^{n}I_{i})=\sum_{i=1}^{n}\mu_{A}(I_{i})

Then μA\mu_{A} is a countably additive \mathcal{E}-valued measure on F()F(\mathbb{R}) as shown in [Lemma 3.7, [Gro14b]] and the functional calculus is defined for elements of 𝕃\mathbb{L} as:

I(f)=i=1naiμA(Si), where f𝕃 as in (1)I(f)=\sum_{i=1}^{n}a_{i}\mu_{A}(S_{i}),\text{ where }f\in\mathbb{L}\text{ as in }(\ref{piecewise_function_definition})

Define 𝕃:={f:¯:(fn)n𝕃, such that fn(t)f(t) for every t}\mathbb{L}^{\uparrow}:=\{f:\mathbb{R}\to\mathbb{\overline{R}}:\exists(f_{n})_{n\in\mathbb{N}}\subset\mathbb{L},\text{ such that }f_{n}(t)\uparrow f(t)\text{ for every }t\in\mathbb{R}\}. Then the integral I:𝕃I:\mathbb{L}\to\mathcal{E} can be extended to an integral from 𝕃\mathbb{L}^{\uparrow} to s\mathcal{E}^{s} as follows: for f𝕃f\in\mathbb{L}^{\uparrow}, we define I(f)=supnI(fn)I(f)=\sup_{n}I(f_{n}) where (fn)(f_{n}) is a sequence in 𝕃\mathbb{L} such that fnff_{n}\uparrow f. Then [Lemma 3.2, [Gro14b]] states that this extension is well-defined. The extension also satisfies the following properties.

Lemma 2 (Lemma 3.4, [Gro14b]).

The extension of II to 𝕃\mathbb{L}^{\uparrow} is well-defined and satisfies the following the properties.

  • If f,g𝕃f,g\in\mathbb{L}^{\uparrow} and fgf\leq g, then I(f)I(g)I(f)\leq I(g);

  • If f𝕃f\in\mathbb{L}^{\uparrow} and 0c<0\leq c<\infty, then cf𝕃cf\in\mathbb{L}^{\uparrow} and I(cf)=cI(f)I(cf)=cI(f);

  • If f,g𝕃f,g\in\mathbb{L}^{\uparrow}, then f+g𝕃f+g\in\mathbb{L}^{\uparrow} and I(f+g)=I(f)+I(g)I(f+g)=I(f)+I(g);

  • If (fn)n(f_{n})_{n\in\mathbb{N}} is a sequence in 𝕃\mathbb{L}^{\uparrow} and fn(t)f(t)f_{n}(t)\uparrow f(t) for every tt , then f𝕃f\in\mathbb{L}^{\uparrow} and I(fn)I(f)I(f_{n})\uparrow I(f).

2.2. Representation of functional calculus

For A¯A\subseteq\mathbb{\overline{R}} and τ:K¯\tau:K\to\overline{\mathbb{R}}, we shall use the notation {τA}\{\tau\in A\} to denote the set {ωK:τ(ω)A}\{\omega\in K:\tau(\omega)\in A\}. We will need the following criterion for the convergence of sequences in a sup-completion. This essentially follows from the results in Section 3 of [BT22]. However they are not stated there in the form of the sup-completion, and thus we repeat certain arguments in the next two lemmas.

Lemma 3.

For G+s,supG=𝟙G\subseteq\mathcal{E}^{s}_{+},\sup G=\infty\mathbbm{1} iff for every non-empty open set UU and every n+n\in\mathbb{R}_{+} there exists a non-empty open set VUV\subseteq U and gGg\in G with g(t)>ng(t)>n for all tVt\in V.

Proof.

Suppose that supG𝟙\sup G\neq\infty\mathbbm{1}. Then there exists f+sf\in\mathcal{E}^{s}_{+} with Gf<𝟙G\leq f<\infty\mathbbm{1}. Hence, there exists a point tKt\in K such that f(t)<f(t)<\infty. Since KK is a totally disconnected space, we can find a clopen subset UKU\subset K containing tt such that f(U)<f(U)<\infty. Hence, by compactness there exists nn\in\mathbb{N} such that ff is less than nn on UU. It follows that every gGg\in G is less than nn on every open subset VV of UU.

Suppose now that there exists an open non-empty set UU and n+n\in\mathbb{R}_{+} such that for every non-empty open subset VUV\subseteq U and gGg\in G there exists tVt\in V such that g(t)ng(t)\leq n. So for a given gGg\in G, we have that {gn}\{g\leq n\} is a closed set and it intersects every open subset of UU. This implies that {gn}\{g\leq n\} contains UU. Thus, we have that gG{gn}U\cap_{g\in G}\{g\leq n\}\supseteq U and clearly, supG𝟙U¯c+n𝟙U¯<𝟙\sup G\leq\infty\mathbbm{1}_{\overline{U}^{c}}+n\mathbbm{1}_{\overline{U}}<\infty\mathbbm{1}. ∎

The above topological property allows us to deduce properties regarding the convergence in sup-completion.

Lemma 4.

Let X+sX\in\mathcal{E}^{s}_{+} and (Xn)n+s(X_{n})_{n\in\mathbb{N}}\subset\mathcal{E}^{s}_{+} such that XnXX_{n}\uparrow X. Then there exists a co-meagre set MKM\subseteq K such that Xn(t)X(t)X_{n}(t)\uparrow X(t) for every tMt\in M.

Proof.

Let us denote U={X<}¯U=\overline{\{X<\infty\}}, which gives that UU is a clopen set and X𝟙UC(K)X\mathbbm{1}_{U}\in C^{\infty}(K). Let us split all the elements as follows: Xn=Xn𝟙U+Xn𝟙UcX_{n}=X_{n}\mathbbm{1}_{U}+X_{n}\mathbbm{1}_{U^{c}}. We have that Xn𝟙UX𝟙UX_{n}\mathbbm{1}_{U}\uparrow X\mathbbm{1}_{U} and hence by [Remark 4.1, [BT22]], the convergence here is pointwise on a co-meagre subset of UU. It remains to show this for X𝟙UcX\mathbbm{1}_{U^{c}}. Note that for this function we have that the value is \infty everywhere.

Let G=(Xn𝟙Uc)n+sG=(X_{n}\mathbbm{1}_{U^{c}})_{n\in\mathbb{N}}\subset\mathcal{E}_{+}^{s}. For nn\in\mathbb{N}, put Vn=gG{g>n}V_{n}=\bigcup_{g\in G}\{g>n\}. Clearly, VnV_{n} is open and VnUcV_{n}\subseteq U^{c}. For every non-empty open set WW, Lemma 3 gives that there exists tWt\in W and gGg\in G such that g(t)>ng(t)>n. Hence tVnt\in V_{n}. That is, for every open set WW, there exists a point tWt\in W such that tVnt\in V_{n}. Hence, VnV_{n} is a a dense subset of UcU^{c}. Let V:=n=1VnV:=\bigcap_{n=1}^{\infty}V_{n}, then VV is co-meagre subset of UcU^{c}. Let tVt\in V, then for every nn\in\mathbb{N} we have tVnt\in V_{n}, and hence supgGg(t)>n\sup_{g\in G}g(t)>n. It follows that supgGg(t)=\sup_{g\in G}g(t)=\infty. Therefore, upon combining the co-meagre subsets of UU and UcU^{c}, we obtain the desired co-meagre subset of KK. ∎

Given a continuous function fC()f\in C(\mathbb{R}) and XC(K)X\in C^{\infty}(K), let UU be the set on which XX is finite. Clearly, UU is open and dense, and the composition of ff and XX is defined, finite, and continuous on UU. Then the composition extends uniquely to a function in C(K)C^{\infty}(K). We shall denote this extended function by fXf\circ X. Also note that by the Maeda-Ogasawara theory, there is a one-to-one correspondence between clopen subsets of KK and bands in \mathcal{E}. For an element YY\in\mathcal{E}, the clopen set corresponding to the band generated by YY is the set {Y0}¯\overline{\{Y\neq 0\}}. Hence the corresponding band projection of EE is PYE=𝟙{Y0}¯P_{Y}E=\mathbbm{1}_{\overline{\{Y\neq 0\}}}.

Lemma 5.

Let f𝕃f\in\mathbb{L} and {=γ0<γ1<<γn<}\{-\infty=\gamma_{0}<\gamma_{1}<\dots<\gamma_{n}<\infty\} be a partition of \mathbb{R} such that f=a𝟙(γn,)+i=1nai𝟙(γi1,γi]f=a_{\infty}\mathbbm{1}_{(\gamma_{n},\infty)}+\sum_{i=1}^{n}a_{i}\mathbbm{1}_{(\gamma_{i-1},\gamma_{i}]}. Then the Daniell integral of ff is

I(f)=a𝟙K{X>b}¯+i=1nai𝟙{X>γi1}¯{X>γi}¯I(f)=a_{\infty}\mathbbm{1}_{K\setminus\overline{\{X>b\}}}+\sum_{i=1}^{n}a_{i}\mathbbm{1}_{\overline{\{X>\gamma_{i-1}\}}\setminus\overline{\{X>\gamma_{i}\}}}
Proof.

Note that since the weak unit EE is fixed, it corresponds to 𝟙\mathbbm{1} in the stone representation. Now, we will first prove the result in the case when f=𝟙Sf=\mathbbm{1}_{S} where S=(a,b]F()S=(a,b]\in F(\mathbb{R}). So,

I(f)=μA(S)=AbAa=(EP(XbE)+E)(EP(XaE)+E)I(f)=\mu_{A}(S)=A_{b}-A_{a}=(E-P_{(X-bE)^{+}}E)-(E-P_{(X-aE)^{+}}E)
=P(XaE)+EP(XbE)+E=P_{(X-aE)^{+}}E-P_{(X-bE)^{+}}E

For (XaE)+(X-aE)^{+}, we have the corresponding band projection P(XaE)+E=𝟙{X>a}¯P_{(X-aE)^{+}}E=\mathbbm{1}_{\overline{\{X>a\}}}. Hence,

I(f)=P(XaE)+EP(XbE)+E=𝟙{X>a}¯{X>b}¯I(f)=P_{(X-aE)^{+}}E-P_{(X-bE)^{+}}E=\mathbbm{1}_{\overline{\{X>a\}}\setminus\overline{\{X>b\}}}

It is easy to see that {X>a}¯{X>b}¯\overline{\{X>a\}}\setminus\overline{\{X>b\}} is clopen. If S=(a,)S=(a,\infty), the above argument can be adapted to give us

I(f)=μA(S)=AAa=supb(EP(XbE)+E)(EP(XaE)+E)I(f)=\mu_{A}(S)=A_{\infty}-A_{a}=\sup_{b\in\mathbb{R}}(E-P_{(X-bE)^{+}}E)-(E-P_{(X-aE)^{+}}E)
=supb(P(XaE)+EP(XbE)+E)=\sup_{b\in\mathbb{R}}(P_{(X-aE)^{+}}E-P_{(X-bE)^{+}}E)
=supb𝟙{X>a}¯{X>b}¯=\sup_{b\in\mathbb{R}}\mathbbm{1}_{\overline{\{X>a\}}\setminus\overline{\{X>b\}}}
=𝟙{X>a}¯=\mathbbm{1}_{\overline{\{X>a\}}}

Similarly, when S=(,b]S=(-\infty,b], we have I(f)=𝟙K{X>b}¯I(f)=\mathbbm{1}_{K\setminus\overline{\{X>b\}}}. Therefore when f𝕃f\in\mathbb{L} is a piece-wise constant function of the form f=a𝟙(γn,)+i=1nai𝟙(γi1,γi]f=a_{\infty}\mathbbm{1}_{(\gamma_{n},\infty)}+\sum_{i=1}^{n}a_{i}\mathbbm{1}_{(\gamma_{i-1},\gamma_{i}]}, we have:

I(f)=a𝟙K{X>b}¯+i=1nai𝟙{X>γi1}¯{X>γi}¯I(f)=a_{\infty}\mathbbm{1}_{K\setminus\overline{\{X>b\}}}+\sum_{i=1}^{n}a_{i}\mathbbm{1}_{\overline{\{X>\gamma_{i-1}\}}\setminus\overline{\{X>\gamma_{i}\}}}

Lemma 6.

Let f𝕃f\in\mathbb{L} such that ff has bounded support. Then for an element XX\in\mathcal{E}, there exists an open dense set HKH\subseteq K such that I(f)(ω)=f(X(ω))I(f)(\omega)=f(X(\omega)) for every ωH\omega\in H.

Proof.

Let us suppose that (a,b](a,b] is the support of ff. Then there exists a partition {a=α0<α1<<αk=b}\{a=\alpha_{0}<\alpha_{1}<\dots<\alpha_{k}=b\} of (a,b](a,b] such that f=j=1kaj𝟙(αj1,αj]f=\sum_{j=1}^{k}a_{j}\mathbbm{1}_{(\alpha_{j-1},\alpha_{j}]}. Then Lemma 5 gives us

I(f)=j=1kaj𝟙{X>αj1}¯{X>αj}¯I(f)=\sum_{j=1}^{k}a_{j}\mathbbm{1}_{\overline{\{X>\alpha_{j-1}\}}\setminus\overline{\{X>\alpha_{j}\}}}

Now set Hj={X>αj}¯{Xαj}H_{j}=\overline{\{X>\alpha_{j}\}}\cap\{X\leq\alpha_{j}\} and N={X=}N=\{X=\infty\}. Now, {X>αj}\{X>\alpha_{j}\} is an open set and {Xαj}\{X\leq\alpha_{j}\} is a closed set of KK satisfying {X>αj}{Xαj}=\{X>\alpha_{j}\}\cap\{X\leq\alpha_{j}\}=\emptyset which implies that HjH_{j} is a closed nowhere dense set. Since each of the sets are closed and nowhere dense, upon setting M=(j=1kHj)NM=\biggl{(}\bigcup_{j=1}^{k}H_{j}\biggr{)}\cup N we have that MM is a closed nowhere dense set and H:=KMH:=K\setminus M is an open dense set. Let ωH\omega\in H. Then we consider two separate cases:

Case 1.

If X(ω)(a,b]X(\omega)\notin(a,b]: Then f(X(ω))=0=I(f)(ω)f(X(\omega))=0=I(f)(\omega).

Case 2.

If X(ω)(a,b]X(\omega)\in(a,b]: Then there exists j0j_{0} such that X(ω)(αj01,αj0]X(\omega)\in(\alpha_{j_{0}-1},\alpha_{j_{0}}]: Then we have that f(X(ω))=aj0f(X(\omega))=a_{j_{0}}. However, we have that ω{X>αj01}¯\omega\in\overline{\{X>\alpha_{j_{0}-1}\}} and ω{Xαj0}\omega\in\{X\leq\alpha_{j_{0}}\}. By definition of the set HH, ω{X>αj0}¯\omega\notin\overline{\{X>\alpha_{j_{0}}\}} and thus, I(f)(ω)=aj0I(f)(\omega)=a_{j_{0}}. Hence, I(f)(ω)=f(X(ω))I(f)(\omega)=f(X(\omega)).

Therefore, we have that I(f)(ω)=f(X(ω))I(f)(\omega)=f(X(\omega)) for every ωH\omega\in H. This implies that fX=I(f)f\circ X=I(f)\in\mathcal{E} when f𝕃f\in\mathbb{L}. ∎

Lemma 7.

Let fC()f\in C(\mathbb{R}) be a positive continuous function with bounded support. Then for an element XX\in\mathcal{E}, we have I(f)=fX+I(f)=f\circ X\in\mathcal{E}_{+}.

Proof.

Because ff has bounded support implies fλ𝟙(a,b]f\leq\lambda\mathbbm{1}_{(a,b]} for some a,ba,b\in\mathbb{R}. Then I(f)λI(𝟙(a,b])I(f)\leq\lambda I(\mathbbm{1}_{(a,b]})\in\mathcal{E}. As \mathcal{E} is order complete, we have I(f)+I(f)\in\mathcal{E}_{+}. Since ff has bounded support, ff is an uniformly continuous function and there exists a sequence (fk)(f_{k}) in 𝕃\mathbb{L} such that fkff_{k}\uparrow\leq f and (fk)(f_{k}) converges to ff uniformly. WLOG, passing to a subsequence, we have 0ffk1k𝟙0\leq f-f_{k}\leq\frac{1}{k}\mathbbm{1}_{\mathbb{R}}. By Lemma 6, fkXf_{k}\circ X\in\mathcal{E} and thus, fXfkX=(ffk)X(1k𝟙)X=1k𝟙Kf\circ X-f_{k}\circ X=(f-f_{k})\circ X\leq(\frac{1}{k}\mathbbm{1}_{\mathbb{R}})\circ X=\frac{1}{k}\mathbbm{1}_{K}. Therefore, fkXf_{k}\circ X converges to fXf\circ X relatively uniformly in \mathcal{E}. Similarly, I(f)I(fk)=I(ffk)I(1k𝟙)=1k𝟙KI(f)-I(f_{k})=I(f-f_{k})\leq I(\frac{1}{k}\mathbbm{1}_{\mathbb{R}})=\frac{1}{k}\mathbbm{1}_{K}. So I(fk)I(f_{k}) converges to I(f)I(f) relatively uniformly in \mathcal{E}. Since I(fk)=fkXI(f_{k})=f_{k}\circ X by Lemma 6, passing to the limit gives, I(f)=fXI(f)=f\circ X.

In view of the above theorem, we will show that for continuous functions the Daniell functional calculus satisfies a nice representation. The idea for the proof of the theorem can be found in Groblers’ paper. Let 𝕃0\mathbb{L}^{\uparrow}_{0} be the set of all real valued positive functions in 𝕃\mathbb{L}^{\uparrow} and

𝕃u={f:f=gh such that g,h𝕃0 and I(g),I(h)u}\mathbb{L}_{u}=\{f:f=g-h\text{ such that }g,h\in\mathbb{L}^{\uparrow}_{0}\text{ and }I(g),I(h)\in\mathcal{E}^{u}\}

Then analogous to the proof of [Proposition 3.5, [Gro14b]] we can show that 𝕃u\mathbb{L}_{u} is a vector space and II has a well-defined extension to 𝕃u\mathbb{L}_{u} by defining for f=gh𝕃uf=g-h\in\mathbb{L}_{u}, I(f)=I(g)I(h)I(f)=I(g)-I(h). The proof also shows that the extension is positive and linear on 𝕃u\mathbb{L}_{u}.

Theorem 8.

Let fC()f\in C(\mathbb{R}). Then for an element XX\in\mathcal{E}, we have I(f)=fXI(f)=f\circ X.

Proof.

Let fC()f\in C(\mathbb{R}) be a positive continuous function. Then we can find a sequence (fn)C()(f_{n})\subset C(\mathbb{R}) of positive continuous functions with bounded support as follows: fn(t)=f(t)f_{n}(t)=f(t) if t[n,n]t\in[-n,n], and fn(t)=0f_{n}(t)=0 if t[n1,n+1]t\notin[-n-1,n+1]. Then fnff_{n}\uparrow f and combining Lemma 2 and Lemma 4 gives that I(fn)I(f)I(f_{n})\uparrow I(f) pointwise on a co-meagre subset of KK. However, we also have that fnX(ω)fX(ω)f_{n}\circ X(\omega)\uparrow f\circ X(\omega) where ω\omega belongs to {X<}\{X<\infty\}, an open dense set. Since Lemma 7 gives that I(fn)=fnXI(f_{n})=f_{n}\circ X, we can conclude that I(f)=fXI(f)=f\circ X on a co-meagre set and thus on a dense set by the Baire category theorem. As ff is a positive continuous function, implies that fX+uf\circ X\in\mathcal{E}_{+}^{u} and therefore fX+sf\circ X\in\mathcal{E}_{+}^{s}. Since I(f)I(f) and fXf\circ X are continuous functions equal on a dense subset of KK, we have I(f)=fXI(f)=f\circ X everywhere on KK. However, fXuf\circ X\in\mathcal{E}^{u} implies that I(f)=fXuI(f)=f\circ X\in\mathcal{E}^{u}. Hence, C()+𝕃uC(\mathbb{R})_{+}\subseteq\mathbb{L}_{u} and thus, C()𝕃uC(\mathbb{R})\subseteq\mathbb{L}_{u}. So given fC()f\in C(\mathbb{R}), applying the preceding argument to f+f^{+} and ff^{-}, we get

I(f)=I(f+)I(f)=f+XfX=fXI(f)=I(f^{+})-I(f^{-})=f^{+}\circ X-f^{-}\circ X=f\circ X

The following are some simple corollaries resulting from the above theorem.

Corollary 9.

Let XX\in\mathcal{E} and fC()f\in C(\mathbb{R}). Then I(f)=f(X)uI(f)=f(X)\in\mathcal{E}^{u}.

Corollary 10.

Let I:𝕃uuI:\mathbb{L}_{u}\to\mathcal{E}^{u} and f,gC()f,g\in C(\mathbb{R}). Then I(fg)=I(f)I(g),I(fg)=I(f)I(g)I(f\vee g)=I(f)\vee I(g),I(f\wedge g)=I(f)\wedge I(g) and I(|f|)=|I(f)|I(|f|)=|I(f)|. That is, the restriction of II to C()C(\mathbb{R}) is a lattice homomorphism.

The following proposition is a stronger statement than [Proposition 4.6, [Gro14b]] and [Proposition 2.8, [AT17]]. Recall the following convergence criterion for elements of C(K)C^{\infty}(K) from [Theorem 3.7, [BT22]] that states that xnuoxx_{n}\xrightarrow{\mathrm{uo}}x if and only if xnx_{n} converges to xx pointwise on a co-meagre set.

Proposition 11.

Let f:f:\mathbb{R}\to\mathbb{R} be a continuous function. If xnuoxx_{n}\xrightarrow{\mathrm{uo}}x in \mathcal{E} then f(xn)uof(x)f(x_{n})\xrightarrow{\mathrm{uo}}f(x) in u\mathcal{E}^{u}.

Proof.

By the convergence criterion, xn(ω)x(ω)x_{n}(\omega)\to x(\omega) for every ωH\omega\in H where HKH\subseteq K is co-meagre set. By continuity of ff this implies that f[xn(ω)]f[x(ω)]f[x_{n}(\omega)]\to f[x(\omega)] for ωH\omega\in H. Therefore, f(xn)uof(x)f(x_{n})\xrightarrow{\mathrm{uo}}f(x). ∎

2.3. Multivariate functional calculus

Since the Daniell functional calculus for continuous functions corresponds to the pointwise composition of functions, we can extend this to the concept of multivariate continuous functions. Given fC(n,)f\in C(\mathbb{R}^{n},\mathbb{R}) and 𝐗=(Xi)i=1n\mathbf{X}=(X_{i})_{i=1}^{n}\subset\mathcal{E} where nn\in\mathbb{N}, let UU be the set on which all the XiX_{i} are finite. Then UU is an open dense set and f(X1,,Xn)f(X_{1},\dots,X_{n}) is a well-defined continuous function on UU. Then denote f(𝐗)f(\mathbf{X}) to be the unique extension of f(X1,,Xn)f(X_{1},\dots,X_{n}) in C(K)C^{\infty}(K). This enables us to prove the multivariate version of Jensen’s inequality. The univariate version of the Jensen’s inequality in the setting of vector lattices was proved in [Theorem 4.4, [Gro14b]]. Let 𝐗=(X1,,Xn)\mathbf{X}=(X_{1},\dots,X_{n}) and for a given conditional expectation operator 𝔽\mathbb{F} on \mathcal{E} let 𝔽𝐗:=(𝔽X1,,𝔽Xn)\mathbb{F}\mathbf{X}:=(\mathbb{F}X_{1},\dots,\mathbb{F}X_{n}).

Theorem 12.

Let fC(n,)f\in C(\mathbb{R}^{n},\mathbb{R}) be a convex function and 𝐗=(Xi)i=1n\mathbf{X}=(X_{i})_{i=1}^{n}\subset\mathcal{E}. Let 𝔽\mathbb{F} be a conditional expectation defined on \mathcal{E}. If f(𝐗)f(\mathbf{X})\in\mathcal{E}, then 𝔽(f(𝐗))f(𝔽𝐗)\mathbb{F}(f(\mathbf{X}))\geq f(\mathbb{F}\mathbf{X}).

Proof.

Since ff is a convex function, it is a fact from analysis that there exists a sequence of affine functions Lm:nL_{m}:\mathbb{R}^{n}\to\mathbb{R} of the form Lm(t)=am,t+bmL_{m}(t)=\langle a_{m},t\rangle+b_{m} for some amn,bma_{m}\in\mathbb{R}^{n},b_{m}\in\mathbb{R} such that for every tnt\in\mathbb{R}^{n}, we have

f(t)=supmLm(t)f(t)=\sup_{m\in\mathbb{N}}L_{m}(t)

Since fLmf\geq L_{m}, we have f(𝐗)Lm(𝐗)f(\mathbf{X})\geq L_{m}(\mathbf{X}) for every mm. Since 𝔽\mathbb{F} is a positive linear projection and f(𝐗)f(\mathbf{X})\in\mathcal{E}, we have

(2) 𝔽(f(𝐗))𝔽(Lm(𝐗))=Lm(𝔽𝐗),m\mathbb{F}(f(\mathbf{X}))\geq\mathbb{F}(L_{m}(\mathbf{X}))=L_{m}(\mathbb{F}\mathbf{X}),\forall m\in\mathbb{N}

For mm\in\mathbb{N}, let Lm=L1LmL_{m}^{\prime}=L_{1}\vee\dots\vee L_{m}. Then LmL_{m}^{\prime} are an increasing sequence of continuous functions such that f(t)=supmLm(t)f(t)=\sup_{m\in\mathbb{N}}L_{m}^{\prime}(t) and thus Lm(𝔽𝐗)f(𝔽𝐗)L_{m}^{\prime}(\mathbb{F}\mathbf{X})\uparrow f(\mathbb{F}\mathbf{X}) by Theorem 2. By Corollary 10 and 2, we have

Lm(𝔽𝐗)=i=1mLi(𝔽𝐗)𝔽(f(𝐗))L_{m}^{\prime}(\mathbb{F}\mathbf{X})=\bigvee_{i=1}^{m}L_{i}(\mathbb{F}\mathbf{X})\leq\mathbb{F}(f(\mathbf{X}))

Thus, it follows that f(𝔽𝐗)𝔽(f(𝐗))f(\mathbb{F}\mathbf{X})\leq\mathbb{F}(f(\mathbf{X})). ∎

Corollary 13.

Let (Xt(i),𝔽t,t)i=1n(X_{t}^{(i)},\mathbb{F}_{t},\mathcal{F}_{t})_{i=1}^{n} be a finite collection of martingales with the filtration (𝔽t)t(\mathbb{F}_{t})_{t\in\mathbb{N}} and let gC(n,)g\in C(\mathbb{R}^{n},\mathbb{R}) be a convex function. If g(𝐗t)=g(Xt(1),,Xt(n))g(\mathbf{X}_{t})=g(X_{t}^{(1)},\dots,X_{t}^{(n)})\in\mathcal{E} for all tt, then (g(𝐗t),𝔽t,t)(g(\mathbf{X}_{t}),\mathbb{F}_{t},\mathcal{F}_{t}) is a sub-martingale.

Proof.

It follows from Jensen’s inequality that for t<st<s, we have

𝔽t[g(𝐗s)]g[𝔽t(𝐗s)]=g(𝐗t)\mathbb{F}_{t}[g(\mathbf{X}_{s})]\geq g[\mathbb{F}_{t}(\mathbf{X}_{s})]=g(\mathbf{X}_{t})

and thus (g(𝐗t),𝔽t,t)(g(\mathbf{X}_{t}),\mathbb{F}_{t},\mathcal{F}_{t}) is a sub-martingale. ∎

3. Discrete stopping times

In the classical measure setting, given a filtered probability space (Ω,,(t)tT,P)(\Omega,\mathcal{F},(\mathcal{F}_{t})_{t\in T},P), (Xt)tT(X_{t})_{t\in T} an adapted stochastic process and τ\tau a stopping time, we define the stopped process as Xtτ(ω)=Xtτ(ω)(ω)X_{t}^{\tau}(\omega)=X_{t\wedge\tau(\omega)}(\omega). The notion for discrete stopping times has been generalized to the theory of vector lattices by Kuo, Labuschagne and Watson in [KLW04a] and further discussed in [KLW04b]. Below we state the definitions for the stopping time and stopped processes in vector lattices. We will only consider the case of discrete filtrations and stochastic processes.

Definition 14.

Let (𝔽i)i(\mathbb{F}_{i})_{i\in\mathbb{N}} be a filtration on \mathcal{E}. A stopping time (Pi)i(P_{i})_{i\in\mathbb{N}} is defined to be an increasing sequence of band projections such that P0=0P_{0}=0 and 𝔽jPi=Pi𝔽j\mathbb{F}_{j}P_{i}=P_{i}\mathbb{F}_{j} whenever iji\leq j.

In particular, each PiP_{i} is order continuous, 0PiI0\leq P_{i}\leq I and (Pi)\mathcal{R}(P_{i}) is a band in \mathcal{E}, hence an order complete sublattice. Note that PnEnP_{n}E\in\mathcal{F}_{n}, because PnE=Pn𝔽nE=𝔽nPnEP_{n}E=P_{n}\mathbb{F}_{n}E=\mathbb{F}_{n}P_{n}E.

The stopping time (Pi)(P_{i}) is said to be bounded if there exists NN so that Pi=IP_{i}=I for all iNi\geq N. For a bounded stopping time P=(Pi)P=(P_{i}) and an adapted stochastic process (Xi,𝔽i)(X_{i},\mathbb{F}_{i}) define the stopped element as XPX_{P} where XP:=i=1(PiPi1)XiX_{P}:=\sum_{i=1}^{\infty}(P_{i}-P_{i-1})X_{i}.

Lemma 15.

Let (Un)n(U_{n})_{n\in\mathbb{N}} be a sequence of pairwise disjoint clopen sets in KK, and set τ=supnn𝟙Un\tau=\sup_{n\in\mathbb{N}}n\mathbbm{1}_{U_{n}}. Then τC(K)+\tau\in C^{\infty}(K)_{+} with {τ=n}=Un,n\{\tau=n\}=U_{n},\forall n\in\mathbb{N} and (τ){0}{}\mathcal{R}(\tau)\subseteq\mathbb{N}\cup\{0\}\cup\{\infty\}.

Proof.

Since (n𝟙Un)n(n\mathbbm{1}_{U_{n}})_{n\in\mathbb{N}} are pairwise disjoint functions in C(K)+C^{\infty}(K)_{+}, we have that τC(K)\tau\in C^{\infty}(K). Since band projections are order continuous,

𝟙Ukτ=𝟙Uk(supnn𝟙Un)=supnn𝟙UnUk=k𝟙Uk.\mathbbm{1}_{U_{k}}\tau=\mathbbm{1}_{U_{k}}(\sup_{n}n\mathbbm{1}_{U_{n}})=\sup_{n}n\mathbbm{1}_{U_{n}\cap U_{k}}=k\mathbbm{1}_{U_{k}}.

Hence, Uk{τ=k}U_{k}\subseteq\{\tau=k\}^{\circ} for all kk\in\mathbb{N}. But let us suppose that n,αV:={τ=n}\exists n\in\mathbb{N},\exists\alpha\in V:=\{\tau=n\}^{\circ} such that αUn\alpha\notin U_{n}. Then consider the function

T=τ𝟙VUnT=\tau-\mathbbm{1}_{V\setminus U_{n}}

Clearly, τ=T\tau=T on (VUn)c(V\setminus U_{n})^{c} and T(VUn)=n1T(V\setminus U_{n})=n-1 and thus T<τT<\tau. However, 𝟙UkT=𝟙Ukτ𝟙(VUn)Uk=k𝟙Uk𝟙(VUn)Uk.\mathbbm{1}_{U_{k}}T=\mathbbm{1}_{U_{k}}\tau-\mathbbm{1}_{(V\setminus U_{n})\cap U_{k}}=k\mathbbm{1}_{U_{k}}-\mathbbm{1}_{(V\setminus U_{n})\cap U_{k}}. Since VUk=V\cap U_{k}=\emptyset, we have 𝟙(VUn)Uk=0\mathbbm{1}_{(V\setminus U_{n})\cap U_{k}}=0 and thus Tk𝟙Uk,kT\geq k\mathbbm{1}_{U_{k}},\forall k\in\mathbb{N}. But this is a contradiction, and thus, we have {τ=n}=Un\{\tau=n\}^{\circ}=U_{n}.

The support of τ\tau is i=1Ui¯\overline{\bigcup_{i=1}^{\infty}U_{i}} and τ(Ui),i\tau(U_{i})\in\mathbb{N},\forall i\in\mathbb{N}. Let ω(i=1Ui)\omega\in\partial(\bigcup_{i=1}^{\infty}U_{i}). Then there exists a sequence (ωn)i=1Ui(\omega_{n})\subset\bigcup_{i=1}^{\infty}U_{i} such that ωnω\omega_{n}\to\omega and the tail of the sequence does not belong to any UiU_{i}. Therefore, WLOG, we can assume that ωni=nUi\omega_{n}\in\bigcup_{i=n}^{\infty}U_{i} and thus τ(ωn)n\tau(\omega_{n})\geq n. Therefore, τ(ω)=\tau(\omega)=\infty and hence (τ){0}{}\mathcal{R}(\tau)\subseteq\mathbb{N}\cup\{0\}\cup\{\infty\}. However, since the range of τ\tau is discrete, for nn\in\mathbb{N} we have {τ=n}={τ<n12}{τ>n+12}\{\tau=n\}=\{\tau<n-\frac{1}{2}\}\cup\{\tau>n+\frac{1}{2}\}. Hence, {τ=n}\{\tau=n\} is an open set and we have {τ=n}=Un\{\tau=n\}=U_{n}. ∎

Maeda-Ogasawara theorem allows to represent the stopping time in terms of continuous functions on the Stone space of \mathcal{E}. Since the (Pn)(P_{n}) are band projections, each of the PnP_{n} correspond to multiplication by a function of the form 𝟙Wn\mathbbm{1}_{W_{n}} where WnW_{n} is a clopen set, with the (Wn)(W_{n}) being an increasing sequence of clopen sets. Let Pn=PnPn1P_{n}^{\prime}=P_{n}-P_{n-1} for n1n\geq 1 and P0=0P_{0}^{\prime}=0. Then each of the PnP_{n}^{\prime} remains a band projection and corresponds to multiplication by a function of the form 𝟙Un\mathbbm{1}_{U_{n}} where Un:=WnWn1U_{n}:=W_{n}\setminus W_{n-1} are pairwise disjoint clopen sets. Let V=K(n=1Un¯)V=K\setminus(\overline{\bigcup_{n=1}^{\infty}U_{n}}) and set τ=𝟙V+supnn𝟙Un\tau=\infty\mathbbm{1}_{V}+\sup_{n}n\mathbbm{1}_{U_{n}}. Lemma 15 gives that supnn𝟙UnC(K)\sup_{n\in\mathbb{N}}n\mathbbm{1}_{U_{n}}\in C^{\infty}(K) and thus τs\tau\in\mathcal{E}^{s} with (τ){}\mathcal{R}(\tau)\subseteq\mathbb{N}\cup\{\infty\}. Furthermore, for nn\in\mathbb{N}, we have {τ=n}=Un\{\tau=n\}=U_{n} is a clopen set and hence

𝟙{τ=n}=PnE=PnEPn1En\mathbbm{1}_{\{\tau=n\}}=P_{n}^{\prime}E=P_{n}E-P_{n-1}E\in\mathcal{F}_{n}

Since 𝟙{τ=n}n\mathbbm{1}_{\{\tau=n\}}\in\mathcal{F}_{n} we also have 𝟙{τn}n\mathbbm{1}_{\{\tau\leq n\}}\in\mathcal{F}_{n}.

In fact, the converse is true as well. That is, if τs\tau\in\mathcal{E}^{s} satisfying (τ)\mathcal{R}(\tau)\subseteq\mathbb{N}\cup\infty and 𝟙{τ=n}n\mathbbm{1}_{\{\tau=n\}}\in\mathcal{F}_{n} for all nn\in\mathbb{N}, then τ\tau corresponds to a stopping time. To see this, define the operator Pn:P_{n}:\mathcal{E}\to\mathcal{E} via Pn(f)=f𝟙{τn}P_{n}(f)=f\cdot\mathbbm{1}_{\{\tau\leq n\}} for ff\in\mathcal{E}. Clearly, (Pn)n(P_{n})_{n\in\mathbb{N}} is an increasing sequence of band projections such that P0=0P_{0}=0. Then upon using the averaging property of conditional expectation operators, we obtain that

𝔽nPn(f)=𝔽nf𝟙{τn}=𝟙{τn}𝔽n(f)=Pn𝔽n(f)\mathbb{F}_{n}P_{n}(f)=\mathbb{F}_{n}f\cdot\mathbbm{1}_{\{\tau\leq n\}}=\mathbbm{1}_{\{\tau\leq n\}}\cdot\mathbb{F}_{n}(f)=P_{n}\mathbb{F}_{n}(f)

Hence, τ\tau corresponds to a stopping time. We summarize the above in the theorem below.

Theorem 16.

Every stopping time corresponds to an element τs\tau\in\mathcal{E}^{s} that satisfies (τ){}\mathcal{R}(\tau)\subseteq\mathbb{N}\cup\{\infty\} and 𝟙{τ=n}n\mathbbm{1}_{\{\tau=n\}}\in\mathcal{F}_{n} for all nn\in\mathbb{N}. The converse is also true.

Representing the stopping times as above is useful in proving their various properties. A variant of Lemma 17 given below has been proven in [Gro11] by Grobler in the case of continuous processes (i.e., when the index is [0,)[0,\infty)). However, the properties regarding the discrete stopping times can not be deduced directly from there. Hence, we explicitly derive the result using the representation of stopping times and stopped processes from Theorem 16.

Lemma 17.

The set of stopping times is closed under the following operations.

  • If σ,τ\sigma,\tau are stopping times then so are στ,στ\sigma\vee\tau,\sigma\wedge\tau and σ+τ\sigma+\tau.

  • If (τn)n(\tau_{n})_{n\in\mathbb{N}} is a sequence of stopping times, then infτn\inf\tau_{n} and supτn\sup\tau_{n} are stopping times. This includes the case where τn\tau_{n} is increasing (or decreasing) to the limit τ\tau.

Proof.

If σ,τ\sigma,\tau are stopping times then {στn}={σn}{τn}\{\sigma\vee\tau\leq n\}=\{\sigma\leq n\}\cap\{\tau\leq n\} and {στn}={σn}{τn}\{\sigma\wedge\tau\leq n\}=\{\sigma\leq n\}\cup\{\tau\leq n\} and thus we have

𝟙{στn}=𝟙{σn}𝟙{τn}n.\mathbbm{1}_{\{\sigma\vee\tau\leq n\}}=\mathbbm{1}_{\{\sigma\leq n\}}\wedge\mathbbm{1}_{\{\tau\leq n\}}\in\mathcal{F}_{n}.
𝟙{στn}=𝟙{σn}𝟙{τn}n.\mathbbm{1}_{\{\sigma\wedge\tau\leq n\}}=\mathbbm{1}_{\{\sigma\leq n\}}\vee\mathbbm{1}_{\{\tau\leq n\}}\in\mathcal{F}_{n}.

Since, στ,στs\sigma\vee\tau,\sigma\wedge\tau\in\mathcal{E}^{s} with (στ),(στ){}\mathcal{R}(\sigma\vee\tau),\mathcal{R}(\sigma\wedge\tau)\subseteq\mathbb{N}\cup\{\infty\}, we have that στ\sigma\vee\tau and στ\sigma\wedge\tau are stopping times. Also {σ+τn}=0sn{σ=s}{τ=ns}\{\sigma+\tau\leq n\}=\bigcup_{0\leq s\leq n}\{\sigma=s\}\cap\{\tau=n-s\} and 𝟙{σ=s},𝟙{τ=ns}n\mathbbm{1}_{\{\sigma=s\}},\mathbbm{1}_{\{\tau=n-s\}}\in\mathcal{F}_{n}. Hence 𝟙{σ+τn}n\mathbbm{1}_{\{\sigma+\tau\leq n\}}\in\mathcal{F}_{n} and, σ+τ\sigma+\tau is a stopping time.

Let us denote supτn=τ\sup\tau_{n}=\tau. WLOG, (τn)(\tau_{n}) is increasing; otherwise, replace τn\tau_{n} with τ1τn\tau_{1}\vee\dots\vee\tau_{n}. We first observe that since τn+s\tau_{n}\in\mathcal{E}_{+}^{s}, τ\tau is well defined and contained in +s\mathcal{E}_{+}^{s}. We claim that (τ){}\mathcal{R}(\tau)\subseteq\mathbb{N}\cup\{\infty\}. Suppose not. Then there exists a point ωK\omega\in K such that τ(ω)=α{}\tau(\omega)=\alpha\notin\mathbb{N}\cup\{\infty\} and hence there exists a clopen set VKV\subseteq K such that τ(V)(α,α+1)\tau(V)\subseteq(\lfloor\alpha\rfloor,\lfloor\alpha\rfloor+1). Let σ=τ𝟙Vc+α𝟙V\sigma=\tau\cdot\mathbbm{1}_{V^{c}}+\lfloor\alpha\rfloor\mathbbm{1}_{V}. Clearly, στn\sigma\geq\tau_{n} for every nn\in\mathbb{N}, which is a contradiction. Therefore, (τ){}\mathcal{R}(\tau)\subseteq\mathbb{N}\cup\{\infty\} and since τ\tau is a continuous function taking discrete values, {τk}\{\tau\leq k\} is a clopen set for kk\in\mathbb{N}. By Lemma 4, there exists a co-meagre set MKM\subseteq K such that τn(ω)τ(ω)\tau_{n}(\omega)\uparrow\tau(\omega) for every ωM\omega\in M. Then for a fixed kk\in\mathbb{N}, we have

{τk}M=n{τnk}M\{\tau\leq k\}\cap M=\bigcap_{n}\{\tau_{n}\leq k\}\cap M

which implies that 𝟙{τk}=𝟙(n{τnk})\mathbbm{1}_{\{\tau\leq k\}}=\mathbbm{1}_{\bigl{(}\bigcap_{n}\{\tau_{n}\leq k\}\bigr{)}^{\circ}} on a co-meagre set.

We claim that 𝟙(n{τnk})=infn𝟙{τnk}\mathbbm{1}_{\bigl{(}\bigcap_{n}\{\tau_{n}\leq k\}\bigr{)}^{\circ}}=\inf_{n}\mathbbm{1}_{\{\tau_{n}\leq k\}} in \mathcal{E}. Since, n{τnk}{τmk}\bigcap_{n}\{\tau_{n}\leq k\}\subseteq\{\tau_{m}\leq k\} for every mm, we have 𝟙(n{τnk})infn𝟙{τnk}\mathbbm{1}_{\bigl{(}\bigcap_{n}\{\tau_{n}\leq k\}\bigr{)}^{\circ}}\leq\inf_{n}\mathbbm{1}_{\{\tau_{n}\leq k\}}. On the other hand, let σ:=infn𝟙{τnk}\sigma:=\inf_{n}\mathbbm{1}_{\{\tau_{n}\leq k\}}. Using a similar argument as above, we can show that (σ){0}{1}\mathcal{R}(\sigma)\subseteq\{0\}\cup\{1\} and therefore, V:=suppσV:=\operatorname{supp}\sigma is a clopen set. Then V{τnk},nVn{τnk}V(n{τnk})V\subseteq\{\tau_{n}\leq k\},\forall n\in\mathbb{N}\implies V\subseteq\bigcap_{n}\{\tau_{n}\leq k\}\implies V\subseteq\biggl{(}\bigcap_{n}\{\tau_{n}\leq k\}\biggr{)}^{\circ}. Hence, σ𝟙(n{τnk})\sigma\leq\mathbbm{1}_{\bigl{(}\bigcap_{n}\{\tau_{n}\leq k\}\bigr{)}^{\circ}} which establishes the claim. Hence, by the Baire category theorem, 𝟙{τk}=infn𝟙{τnk}k\mathbbm{1}_{\{\tau\leq k\}}=\inf_{n}\mathbbm{1}_{\{\tau_{n}\leq k\}}\in\mathcal{F}_{k} and thus supnτn\sup_{n}\tau_{n} is a stopping time. Similarly, we can argue for infτn\inf\tau_{n}. ∎

Remark 18.

Let τ\tau be a stopping time and (nk)(n_{k}) be an increasing sequence in \mathbb{N}. Then clearly there exists a positive, increasing continuous function g:+{}+{}g:\mathbb{R}_{+}\cup\{\infty\}\to\mathbb{R}_{+}\cup\{\infty\} such that g(k)=nk,g(0)=0,g()=g(k)=n_{k},g(0)=0,g(\infty)=\infty and g(t)tg(t)\geq t for all tt. Since gg is a continuous function, g(τ)g(\tau) is a well defined element of +s\mathcal{E}^{s}_{+} with (g(τ))={}\mathcal{R}(g(\tau))=\mathbb{N}\cup\{\infty\}. We claim that g(τ)g(\tau) is again a stopping time. Now, {g(τ)=nk}={τ=k}\{g(\tau)=n_{k}\}=\{\tau=k\} and hence 𝟙{g(τ)=nk}=𝟙{τ=k}knk\mathbbm{1}_{\{g(\tau)=n_{k}\}}=\mathbbm{1}_{\{\tau=k\}}\in\mathcal{F}_{k}\subseteq\mathcal{F}_{n_{k}}. If n(nk)k=1n\notin(n_{k})_{k=1}^{\infty} then {g(τ)=n}=\{g(\tau)=n\}=\emptyset and 𝟙{g(τ)=n}=0n\mathbbm{1}_{\{g(\tau)=n\}}=0\in\mathcal{F}_{n}. Thus, g(τ)g(\tau) is a stopping time. We note that g(τ)g(\tau) is only determined by τ\tau and (nk)(n_{k}), and does not depend on the choice of gg.

For a given stopping time τs\tau\in\mathcal{E}^{s} and nn\in\mathbb{N}, the above theorem shows that τn𝟙\tau\wedge n\mathbbm{1} is also a stopping time. So for an adapted process (Xn)n(X_{n})_{n\in\mathbb{N}}, there exists a stopped element Xτn𝟙X_{\tau\wedge n\mathbbm{1}} for every nn\in\mathbb{N}. So we define the stopped process corresponding to τ\tau and (Xn)n(X_{n})_{n\in\mathbb{N}} to be the sequence of stopped elements (Xτn𝟙)n(X_{\tau\wedge n\mathbbm{1}})_{n\in\mathbb{N}}.

It is evident from above that a stopping time τ\tau is bounded precisely when τC(K)\tau\in C(K). The above representation of stopping times enables us to do the same for the stopped processes. Let τ\tau be a bounded stopping time. Then there exists NN\in\mathbb{N} such that (τ)N\mathcal{R}(\tau)\leq N. Since PnPn1=PnP_{n}-P_{n-1}=P_{n}^{\prime}, given an adapted process (Xn,𝔽n)n(X_{n},\mathbb{F}_{n})_{n\in\mathbb{N}} we have PnXn=Xn.1{τ=n}P_{n}^{\prime}X_{n}=X_{n}.\mathbbm{1}_{\{\tau=n\}}. Hence, Xτ=n=1Xn.1{τ=n}X_{\tau}=\sum_{n=1}^{\infty}X_{n}.\mathbbm{1}_{\{\tau=n\}}. Moreover, as τ\tau is a continuous function with values in \mathbb{N}, this implies that {τ=n}\{\tau=n\} is a clopen set for every nNn\leq N and {τ=n}=\{\tau=n\}=\emptyset when n>Nn>N. Therefore, Xτ=n=1NXn𝟙{τ=n}X_{\tau}=\sum_{n=1}^{N}X_{n}\cdot\mathbbm{1}_{\{\tau=n\}}. Let ωK\omega\in K, then evaluated pointwise, we have Xτ(ω)=Xτ(ω)(ω)X_{\tau}(\omega)=X_{\tau(\omega)}(\omega). Thus, the stopped process is (Xτn𝟙)n(X_{\tau\wedge n\mathbbm{1}})_{n\in\mathbb{N}} evaluated pointwise.

Theorem 19.

Let (Xn,𝔽n)n(X_{n},\mathbb{F}_{n})_{n\in\mathbb{N}} be an adapted process on \mathcal{E}. Then, given a bounded discrete stopping time τ\tau, the stopped element XτX_{\tau}\in\mathcal{E} is the function evaluated pointwise. That is, at ωK\omega\in K, the value of the stopped element is Xτ(ω)(ω)X_{\tau(\omega)}(\omega).

The following proposition improves the result [Lemma 5.3, [KLW04a]].

Proposition 20.

Let (Xn,𝔽n)n(X_{n},\mathbb{F}_{n})_{n\in\mathbb{N}} be an increasing adapted process. Then the following statements hold.

  • Let σ\sigma and τ\tau be two bounded stopping times. Then Xστ=XσXτX_{\sigma\vee\tau}=X_{\sigma}\vee X_{\tau} and Xστ=XσXτX_{\sigma\wedge\tau}=X_{\sigma}\wedge X_{\tau}.

  • Let τn\tau_{n} be a sequence of bounded stopping times such that τ:=supnτn\tau:=\sup_{n}\tau_{n} is bounded. Then Xτ=supnXτnX_{\tau}=\sup_{n}X_{\tau_{n}}.

  • Let τn\tau_{n} be a sequence of bounded stopping times. Then Xinfnτn=infnXτnX_{\inf_{n}\tau_{n}}=\inf_{n}X_{\tau_{n}}.

Proof.

By Lemma 17, we have that στ\sigma\vee\tau is a stopping time, and therefore, XστX_{\sigma\vee\tau} is a continuous function in \mathcal{E}. Let ωK\omega\in K, then X(στ)(ω)(ω)=Xσ(ω)τ(ω)(ω)X_{(\sigma\vee\tau)(\omega)}(\omega)=X_{\sigma(\omega)\vee\tau(\omega)}(\omega). WLOG, τ(ω)σ(ω)\tau(\omega)\geq\sigma(\omega). Then X(στ)(ω)(ω)=Xτ(ω)(ω)X_{(\sigma\vee\tau)(\omega)}(\omega)=X_{\tau(\omega)}(\omega). Since (Xn)(X_{n}) is an increasing process, we have Xτ(ω)Xσ(ω)X_{\tau(\omega)}\geq X_{\sigma(\omega)}. Therefore, we have: X(στ)(ω)(ω)=Xτ(ω)(ω)Xσ(ω)(ω)X_{(\sigma\vee\tau)(\omega)}(\omega)=X_{\tau(\omega)}(\omega)\vee X_{\sigma(\omega)}(\omega). Since the equality is valid for every point in KK, we have Xστ=XσXτX_{\sigma\vee\tau}=X_{\sigma}\vee X_{\tau}. Similarly, we can conclude for Xστ=XσXτX_{\sigma\wedge\tau}=X_{\sigma}\wedge X_{\tau}.

WLOG, (τn)(\tau_{n}) is increasing; otherwise, replace τn\tau_{n} with τ1τn\tau_{1}\vee\dots\vee\tau_{n}. Since τn\tau_{n} and τ\tau are bounded stopping times, XτnX_{\tau_{n}} and XτX_{\tau} are well defined elements of \mathcal{E}. For an arbitrary ωK\omega\in K, we have Xτ(ω)(ω)Xτn(ω)(ω)X_{\tau(\omega)}(\omega)\geq X_{\tau_{n}(\omega)}(\omega) for every nn\in\mathbb{N}. Therefore, XτsupXτnX_{\tau}\geq\sup X_{\tau_{n}}. Let σXτn\sigma\geq X_{\tau_{n}} for every nn\in\mathbb{N}. By [Lemma 3.6, [BT22]], there exists a co-meagre subset DKD\subseteq K such that τn(ω)τ(ω)\tau_{n}(\omega)\uparrow\tau(\omega) for every ωD\omega\in D. Fix ωD\omega\in D. Since τ\tau is bounded, for large enough nn we have τn(ω)=τ(ω)\tau_{n}(\omega)=\tau(\omega). Hence, σ(ω)Xτn(ω)(ω)=Xτ(ω)(ω)\sigma(\omega)\geq X_{\tau_{n}(\omega)}(\omega)=X_{\tau(\omega)}(\omega). Therefore, by the Baire Category theorem Xτ=supnXτnX_{\tau}=\sup_{n}X_{\tau_{n}}. Similarly, we can prove for the infimum. ∎

The following theorem is a vector lattice version of the Début theorem [Fis13] for discrete stochastic processes. For any stopping time, there exists an adapted stochastic process and a subset of \mathbb{R} such that the corresponding hitting time will be precisely this stopping time. The stochastic process can be chosen intuitively and similar to the classical probability case. It will be 11 until just before the stopping time is reached, from which on, it will be 0. The increasing process, therefore, first hits the set {1}\{1\} at the stopping time.

Theorem 21.

Let τ\tau be a discrete stopping time. Then there exists an adapted process (Xn)n(X_{n})_{n\in\mathbb{N}} such that τ(ω)=inf{t:Xt(ω)=1}\tau(\omega)=\inf\{t\in\mathbb{N}:X_{t}(\omega)=1\}.

Proof.

Let Sn={τn}S_{n}=\{\tau\leq n\} and let Xn=𝟙SnX_{n}=\mathbbm{1}_{S_{n}} for all nn\in\mathbb{N}. Since 𝟙{τn}n\mathbbm{1}_{\{\tau\leq n\}}\in\mathcal{F}_{n}, this implies that XnnX_{n}\in\mathcal{F}_{n} and thus, the stochastic process (Xn)n(X_{n})_{n\in\mathbb{N}} is increasing and adapted. Let σ:K¯\sigma:K\to\overline{\mathbb{R}} defined via σ(ω)=inf{t:Xt(ω)=1}\sigma(\omega)=\inf\{t\in\mathbb{N}:X_{t}(\omega)=1\}. Then {σ=n}={Xn1=0}{Xn=1}\{\sigma=n\}=\{X_{n-1}=0\}\cap\{X_{n}=1\}. However, we also have {τ=n}={Xn1=0}{Xn=1}\{\tau=n\}=\{X_{n-1}=0\}\cap\{X_{n}=1\} for every nn\in\mathbb{N}. Moreover, ω{τ=}Xn(ω)=0,nσ(ω)=inf{}=\omega\in\{\tau=\infty\}\iff X_{n}(\omega)=0,\forall n\in\mathbb{N}\iff\sigma(\omega)=\inf\{\emptyset\}=\infty. Hence, τ(ω)=inf{t:Xt(ω)=1}\tau(\omega)=\inf\{t\in\mathbb{N}:X_{t}(\omega)=1\}. ∎

Remark 22.

There is also the notion of stopping times for continuous stochastic processes in vector lattices, as discussed by Grobler in [Gro10, Gro11, Gro21]. Let T+T\subset\mathbb{R}_{+} be an interval. When phrased in terms of the C(K)C^{\infty}(K) representation, the stopping time for the filtration (𝔽t)tT(\mathbb{F}_{t})_{t\in T} is an orthomorphism 𝕊Orth()+\mathbb{S}\in\text{Orth}(\mathcal{E})_{+} such that 𝟙{𝕊t}𝔉t\mathbbm{1}_{\{\mathbb{S}\leq t\}^{\circ}}\in\mathfrak{F}_{t} for every tTt\in T. However, we can not emulate the above results regarding stopped processes for the continuous case. This is because an arbitrary union of nowhere dense sets will need not be nowhere dense. Moreover, we use the Baire category theorem at various points in this paper, which does not hold for arbitrary union of nowhere dense sets.

Acknowledgment

I would like to thank my advisor, Prof. Vladimir Troitsky, for his patient guidance, from the construction of the general idea and the argument to the details of writing this paper.

References

  • [AB03] Charalambos D. Aliprantis and Owen Burkinshaw. Locally solid Riesz spaces with applications to economics, volume 105 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, second edition, 2003.
  • [AB06] Charalambos D. Aliprantis and Owen Burkinshaw. Positive operators. Springer, Dordrecht, 2006. Reprint of the 1985 original.
  • [AN] Youssef Azouzi and Youssef Nasri. Generating functions in riesz spaces. preprint. arXiv:2212.00673 [math.FA].
  • [AN22] Youssef Azouzi and Youssef Nasri. The sup-completion of a Dedekind complete vector lattice. J. Math. Anal. Appl., 506(2):Paper No. 125651, 20, 2022.
  • [AR18] Youssef Azouzi and Kawtar Ramdane. On the distribution function with respect to conditional expectation on Riesz spaces. Quaest. Math., 41(2):257–264, 2018.
  • [AT17] Youssef Azouzi and Mejdi Trabelsi. LpL^{p}-spaces with respect to conditional expectation on Riesz spaces. J. Math. Anal. Appl., 447(2):798–816, 2017.
  • [Azo19] Youssef Azouzi. Completeness for vector lattices. J. Math. Anal. Appl., 472(1):216–230, 2019.
  • [BT22] Eugene Bilokopytov and Vladimir G. Troitsky. Order and uo-convergence in spaces of continuous functions. Topology Appl., 308:Paper No. 107999, 9, 2022.
  • [Don82] Klaus Donner. Extension of positive operators and Korovkin theorems, volume 904 of Lecture Notes in Mathematics. Springer-Verlag, Berlin-New York, 1982.
  • [Fis13] Tom Fischer. On simple representations of stopping times and stopping time sigma-algebras. Statist. Probab. Lett., 83(1):345–349, 2013.
  • [Gro10] J. J. Grobler. Continuous stochastic processes in Riesz spaces: the Doob-Meyer decomposition. Positivity, 14(4):731–751, 2010.
  • [Gro11] J. J. Grobler. Doob’s optional sampling theorem in Riesz spaces. Positivity, 15(4):617–637, 2011.
  • [Gro14a] J. J. Grobler. The Kolmogorov-Čentsov theorem and Brownian motion in vector lattices. J. Math. Anal. Appl., 410(2):891–901, 2014.
  • [Gro14b] Jacobus Grobler. Jensen’s and martingale inequalities in Riesz spaces. Indag. Math. (N.S.), 25(2):275–295, 2014.
  • [Gro21] Jacobus J. Grobler. Stopped processes and Doob’s optional sampling theorem. J. Math. Anal. Appl., 497(1):Paper No. 124875, 13, 2021.
  • [Kal17] Olav Kallenberg. Random measures, theory and applications, volume 77 of Probability Theory and Stochastic Modelling. Springer, Cham, 2017.
  • [KLW04a] Wen-Chi Kuo, Coenraad C. A. Labuschagne, and Bruce A. Watson. Discrete-time stochastic processes on Riesz spaces. Indag. Math. (N.S.), 15(3):435–451, 2004.
  • [KLW04b] Wen-Chi Kuo, Coenraad C. A. Labuschagne, and Bruce A. Watson. Riesz space and fuzzy upcrossing theorems. In Soft methodology and random information systems, Adv. Soft Comput., pages 101–108. Springer, Berlin, 2004.
  • [KLW05] Wen-Chi Kuo, Coenraad C. A. Labuschagne, and Bruce A. Watson. Conditional expectations on Riesz spaces. J. Math. Anal. Appl., 303(2):509–521, 2005.
  • [PT] Achintya Raya Polavarapu and Vladimir G. Troitsky. A representation of sup-completion. preprint. arXiv:2306.06248 [math.FA].