Skip to main content
Log in

No-arbitrage conditions and pricing from discrete-time to continuous-time strategies

  • Research Article
  • Published:
Annals of Finance Aims and scope Submit manuscript

Abstract

In this paper, a general framework is developed for continuous-time financial market models defined from simple strategies through conditional topologies that avoid stochastic calculus and do not necessitate semimartingale models. We then compare the usual no-arbitrage conditions of the literature, e.g. the usual no-arbitrage conditions NFL, NFLVR and NUPBR and the recent AIP condition. With appropriate pseudo-distance topologies, we show that they hold in continuous time if and only if they hold in discrete time. Moreover, the super-hedging prices in continuous time coincide with the discrete-time super-hedging prices, even without any no-arbitrage condition.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

References

  • Baptiste, J., Carassus, L., Lépinette, E.: Pricing without martingale measure. Preprint, https://hal.archives-ouvertes.fr/hal-01774150, (2020)

  • Barron, E.N., Cardaliaguet, P., Jensen, R.: Conditional essential suprema with applications. Appl Math Optim 48, 229–253 (2003)

    Article  Google Scholar 

  • Bayraktar, E., Sayit, H.: No arbitrage conditions for simple trading strategies. Ann Finance 6(1), 147–156 (2010)

    Article  Google Scholar 

  • Bender, C., Sottinen, T., Valkeila, E.: Pricing by hedging and no-arbitrage beyond semi-martingale. Finance Stochast 12(4), 441–468 (2008)

    Article  Google Scholar 

  • Carassus, L., Lépinette, E.: Pricing without no-arbitrage condition in discrete-time. J Math Anal Appl 505(1), 125441 (2021)

    Article  Google Scholar 

  • Cheredito, P.: Arbitrage in fractional Brownian motion models. Finance Stochast 7(4), 533–553 (2003)

    Article  Google Scholar 

  • Dalang, E.C., Morton, A., Willinger, W.: Equivalent martingale measures and no-arbitrage in stochastic securities market models. Stochast Stochast Rep 29, 185–201 (1990)

    Article  Google Scholar 

  • Delbaen, F., Schachermayer, W.: A general version of the fundamental theorem of asset pricing. Math Ann 300, 463–520 (1994)

    Article  Google Scholar 

  • El Mansour, M., Lépinette, E.: Conditional interior and conditional closure of a random sets. J Optim Theory Appl 187, 356–369 (2020)

    Article  Google Scholar 

  • Filipovic, D., Kupper, M., Vogelpoth, N.: Separation and duality in locally \(L^0\) convex modules. J Funct Anal 256(12), 3996–4029 (2009)

    Article  Google Scholar 

  • Guasoni, P.: No arbitrage with transaction costs, with fractional Brownian motion and beyond. Math Financ 16(2), 469–588 (2006)

    Google Scholar 

  • Jacod, J., Shiryaev, A.N.: Limit theorems for stochastic processes. Grundlehren der mathematischen Wissenschaften 288, a series of comprehensive studies in mathematics. Springer: Berlin (2003)

  • Jarrow, R., Protter, P., Sayit, H.: No arbitrage without semimartingales. Ann Appl Probab 19(2), 596–616 (2009)

    Article  Google Scholar 

  • Kabanov, Y., Stricker, C.: A Teachers’ note on no-arbitrage criteria. In: Séminaire de Probabilités, XXXV, volume 1755 of Lecture Notes in Mathematics, Springer: Berlin, 149–152 (2001)

  • Kabanov, Y., Safarian, M.: Markets with Transaction Costs. Mathematical Theory. Springer, Berlin (2009)

    Google Scholar 

  • Karatzas, I., Kardaras, C.: The numéraire portfolio in semimartingale financial models. Finance Stochast 11, 447–493 (2007)

    Article  Google Scholar 

  • Kreps, D.M.: Arbitrage and equilibrium in economies with infinitely many commodities. J Math Econ 8, 15–35 (1981)

    Article  Google Scholar 

  • Lépinette, E., Molchanov, I.: Conditional cores and conditional convex hulls of random sets. J Math Anal Appl 478(2), 368–392 (2019)

    Article  Google Scholar 

  • Lo, A.W.: Long-term memory in stock market prices. Econometrica 59(5), 1279–1313 (1991)

    Article  Google Scholar 

  • Pakkanen, M.: Stochastic integrals and conditional full support. J Appl Probab 47(3), 650–667 (2010)

    Article  Google Scholar 

  • Rockafellar, R.T., Wets, R.J.B.: Variational Analysis, Grundlehren der mathematischen Wissenschaften, vol. 317. Springer, Berlin (1998)

    Google Scholar 

  • Rogers, L.C.G.: Arbitrage with fractional Brownian motion. Math Financ 7, 95–105 (1997)

    Article  Google Scholar 

  • Sayit, H.: Absence of arbitrage in a general framework. Ann Finance 9, 611–624 (2013)

    Article  Google Scholar 

  • Sottinen, T.: Fractional Brownian motion, random walks and binary market models. Finance Stochast 5(3), 343–355 (2001)

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Emmanuel Lépinette.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Dorsaf Cherif and Emmanuel Lépinette contributed equally to this work.

Appendices

Appendix A: Proof of Proposition 10

The proof of Proposition 10 is deduced from Lemma 37. To get it, we first show intermediate steps such as the following Lemmas 33,  34,  35 and  36.

Lemma 33

Let \((M_t)_{t\in [0,T]}\) be a sub-maxingale. Let \( \tau \) be a stopping time such that \(\tau (\Omega )= \{ t_1, t_2, \cdots ,t_n \}\) where \((t_i)_{i=1}^n\) is an increasing sequence of discrete dates. Then, for all \(i=1,\cdots ,n\), we have \( {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{t_i}}(M_{\tau }) \ge M_{\tau \wedge t_i}\).

Proof

We have:

$$\begin{aligned} {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{t_i}}(M_{\tau \wedge t_{i+1}})= & {} {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{t_i}}(M_{\tau \wedge t_{i+1}} 1_{\{\tau \le t_i \}}) +{{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{t_i}}(M_{\tau \wedge t_{i+1}} 1_{\{\tau>t_i \}}),\\= & {} 1_{\{\tau>t_i \}}{{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{t_i}}(M_{ t_{i+1}}) + 1_{\{\tau \le t_i \}} {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{t_i}}(M_{\tau \wedge t_{i}} ),\\\ge & {} 1_{\{\tau >t_i \}}M_{ t_{i}} + 1_{\{\tau \le t_i \}} M_{\tau \wedge t_{i}}= M_{\tau \wedge t_i}. \end{aligned}$$

If \(j >i+1\), argue by induction. By the tower property, we first have \({{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{t_i}}(M_{\tau \wedge t_{j}}) ={{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{t_i}}({{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{t_{j-1}}}(M_{\tau \wedge t_{j}}))\). Therefore, by the first step above, \({{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{t_i}}(M_{\tau \wedge t_{j}})\ge {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{t_i}}(M_{\tau \wedge t_{j-1}})\) and we conclude by induction. \(\square \)

Lemma 34

Let \( \tau \) be a stopping time such that \(\tau (\Omega )= \{ t_1, t_2, \cdots ,t_n \}\) where \((t_i)_{i=1}^n\) is an increasing sequence of discrete dates. Then, for any random variable X, we have

$$\begin{aligned} {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau }}(X 1_{\{\tau = t_i\}})= {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{t_i}} (X)1_{\{\tau = t_i\}}. \end{aligned}$$

Proof

As \(1_{\{\tau = t_i\}}\) is \({\mathcal {F}}_{\tau }\)-mesurable, then we get that

$$\begin{aligned} {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau }}(X1_{\{\tau = t_i\}})={{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau }}(X) 1_{\{\tau = t_i\}}. \end{aligned}$$

Since \(X 1_{\{\tau = t_i\}}\le {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{t_i}}(X )1_{\{\tau =t_i\}}\), we deduce that

$$\begin{aligned} {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau }}(X 1_{\{\tau = t_i\}})\le \ {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau }} ({{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{t_i}}(X )1_{\{\tau = t_i\}}). \end{aligned}$$

We claim that \(Z={{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{t_i}}(X )1_{\{\tau = t_i\}}\) is \({\mathcal {F}}_{\tau }\)-mesurable. For any \(k \in \textbf{R}\),

$$\begin{aligned} \{Z \le k \} = \{0 \le k \} \cap \{ \tau \ne t_i \} \cup \{ \tau = t_i \} \cap \{ {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{t_i}}(X) \le k \}. \end{aligned}$$

Note that \(\{0 \le k \} = \emptyset \) or \(\Omega \) and \(\{ \tau \ne t_i \} \in {\mathcal {F}}_{\tau }\) hence \(\{0 \le k \} \cap \{ \tau \ne t_i \}\in {\mathcal {F}}_{\tau }\). Now let us show that \(B = \{ \tau = t_i \} \cap \{ {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{t_i}}(X) \le k \} \in {\mathcal {F}}_{\tau }\). To do so, we evaluate \(B \cap \{ \tau \le t \}\) for \(t\ge 0\). Note that \(t_j \le t < t_{j+1}\) for some \(t_j \in \{t_0,\cdots ,t_n,t_{n+1}\}\), where \(t_{n+1}=\infty \). So, we deduce that \(B \cap \{ \tau \le t \}\) coincides with \(B \cap \{ \tau \le t_j \}= \emptyset \) if \(t_j < t_i\). Otherwise, we obtain that \(B \cap \{ \tau \le t \}=B \in F_{t_i} \subseteq F_{t_j} \subseteq F_t.\) Therefore, \(B \cap \{ \tau \le t \}\in {\mathcal {F}}_t\), for all \(t\in \textbf{R}\), hence \(B \in {\mathcal {F}}_{\tau }\). Finally, Z is \({\mathcal {F}}_{\tau }\)-mesurable and the inequality \({{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau }}(X 1_{\{\tau = t_i\}})\le \ {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{t_i}}(X )1_{\{\tau = t_i\}} \) holds. For the reverse inequality it suffices to show that \(Y={{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau }}(X 1_{\{\tau = t_i\}} )\) is \({\mathcal {F}}_{t_i}\)-measurable. Since \(\{\tau \ne t_i\}\in {\mathcal {F}}_{\tau }\), we get that \(Y 1_{\{\tau \ne t_i\}}=0\) and

$$\begin{aligned} \{ Y \le k \} = (\{0 \le k \} \cap \{\tau \ne t_i \}) \cup (A \cap \{\tau = t_i \}), \end{aligned}$$

with \(A=\{{{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau }}(X 1_{\{\tau = t_i\}}) \le k \} \). As \(A \in {\mathcal {F}}_{\tau }\), \( A \cap \{\tau \le t_i \} \in {\mathcal {F}}_{t_i}\) and, finally, \(A \cap \{\tau = t_i \}= A \cap \{\tau = t_i \} \cap \{\tau \le t_i \}\in {\mathcal {F}}_{t_i}\). Therefore, for all \(k\in \textbf{R}\), \( \{ Y \le k \}\in {\mathcal {F}}_{t_i}\), i.e. Y is \({\mathcal {F}}_{t_i}\)-measurable. At last, notice that \({{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau }}(X 1_{\{\tau = t_i\}} )\ge X 1_{\{\tau = t_i\}} \) and, since Y is \({\mathcal {F}}_{t_i}\)-measurable, we get that \({{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau }}(X 1_{\{\tau = t_i\}} )\ge {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{t_i}}(X 1_{\{\tau = t_i\}})\). The conclusion follows. \(\square \)

Lemma 35

Let \((M_t)_{t\in [0,T]}\) be a sub-maxingale. Let \( \tau \), S be two stopping times. Suppose that \(S(\Omega )= \{ t_1, t_2, \cdots ,t_n \}\) where \((t_i)_{i=1}^n\) is an increasing sequence of discrete dates and suppose that \(\tau (\Omega )\) is also a finite set. Then \({{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{S}}(M_{\tau })\ge M_{\tau \wedge s}\).

Proof

By Lemma 34, we obtain \({{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{S}}(M_{\tau }) =\sum \nolimits _{i=1}^n {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{t_i}}(M_{\tau }) 1_{\{S=t_i\}}\). By Lemma 33, we deduce that

$$\begin{aligned} {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{S}}(M_{\tau }) \ge \sum _{i=1}^n M_{\tau \wedge t_i} 1_{\{S=t_i\}}= \sum _{i=1}^n M_{ \tau \wedge S} 1_{\{S=t_i\}} = M_{ \tau \wedge S}. \end{aligned}$$

\(\square \)

Lemma 36

Let \(\tau \in [0,T]\) be a stopping time. Suppose that the filtration \(({\mathcal {F}}_t)_{t\in [0,T]}\) is right-continuous. There exists a non increasing sequence \((\tau _n)_n\) of stopping times converging to \(\tau \) such that, for any \(X\in L^0(\textbf{R},{\mathcal {F}}_T)\),

$$\begin{aligned} {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau }}(X)= \lim _n \uparrow {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau _n}}(X). \end{aligned}$$

Moreover, \(\tau ^n(\Omega )\) is finite for all \(n\ge 1\).

Proof

Let \(\tau \) be a stopping time taking values in [0, T]. For any \(n\ge 1\), we define \(\tau ^n(\omega ) = T(i+1)/2^n\) where \(i=i(\omega )\) is uniquely defined such that \(Ti/2^n< \tau (\omega ) \le T(i+1)/2^n\) for \(i\ge 1\) or \(0\le \tau (\omega ) \le T/2^n\) when \(i=0\). Note that \(\tau ^n(\Omega )\) is finite and \(\tau ^n \ge \tau \). It is easily seen that \((\tau _n)_n\) is non increasing, positive and \(\lim _n \tau _n=\tau \). Moreover, \(\tau ^n\) is a stopping time. Indeed, for any fixed \(t \in [0,T)\), there exists \(i \in {\mathbb {N}}\) such that \(Ti/2^n\le t < T(i+1)/2^n\). Then \(\{ \tau ^n \le t \}= \{ \tau \le Ti/2^n \} \in {\mathcal {F}}_{Ti/2^n} \subset {\mathcal {F}}_t\) and the conclusion follows.

As \((\tau ^n)_n\) is non increasing, then \(({\mathcal {F}}_{\tau ^n})_n\) non increasing. As we know that \({{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau ^{n+1}}}(X) \ge X\) and \({{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau ^{n+1}}}(X) \) is \({\mathcal {F}}_{\tau ^n}\)-measurable (\(\tau _{n+1}\le \tau _n)\), we deduce that \({{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau ^n}}(X) \le {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau ^{(n+1)}}}(X) \), i.e. \(({{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau ^n}}(X))_n\) is non decreasing.

Similarly, \(\tau ^n \ge \tau \) implies that \({{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau ^n}}(X) \le {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau }}(X) \). Therefore, \(\lim _n \uparrow {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau ^n}}(X) \le {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau }}(X) \). To obtain the reverse inequality, we consider the sequence \(({{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau + T/n}}(X))_n.\) Since \(\tau + T/n \ge \tau ^n\), then

$$\begin{aligned} \lim _n \uparrow {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau + T/n}}(X) \le \lim \uparrow {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau ^n}}(X) \le {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau }}(X). \end{aligned}$$

It suffices to see that \(Z=\lim _n \uparrow {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau +T/n}}(X)\) is \({\mathcal {F}}_{\tau }\)-measurable to conclude. Indeed, \(Z\ge X\) hence \(Z\ge {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau }}(X) \) and inequalities above are equalities. For all \(k\in \textbf{R}\), \(t\ge 0\), and any \(n_0\ge 1\)

$$\begin{aligned} \{Z\le k\}\cap \{\tau \le t\}= & {} \bigcap _{n\ge 1} \{{{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau + T/n}}(X)\le k\}\cap \{\tau \le t\},\\= & {} \bigcap _{n\ge n_0}\{ {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau + T/n}}(X)\le k\} \cap \{\tau + T/n\le t+ T/n\}. \end{aligned}$$

Notice that \({{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau + T/n}}(X)\) is \({\mathcal {F}}_{\tau +T/n}\)-measurable. We deduce that:

$$\begin{aligned}{} & {} \{{{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau + T/n}}(X)\le k\}\in {\mathcal {F}}_{\tau +T/n},\\{} & {} \{{{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{\tau + T/n}}(X)\le k\}\cap \{\tau + T/n \le t+ T/n\}\in {\mathcal {F}}_{t+ T/n}. \end{aligned}$$

Therefore, for any \(\epsilon >0\) and \(n_0\ge 1\) such that \(t+ T/n\le t+\epsilon \), we have \({\mathcal {F}}_{t+ T/n}\subseteq {\mathcal {F}}_{t+\epsilon }\) and, finally, \(\{Z\le k\}\cap \{\tau \le t\}\in \cap _{\epsilon >0}{\mathcal {F}}_{t+\epsilon }={\mathcal {F}}_{t+}={\mathcal {F}}_t\). We deduce that \(\{Z\le k\}\in {\mathcal {F}}_{\tau }\), for all \(k\in \textbf{R}\), i.e. Z is \({\mathcal {F}}_{\tau }\)-measurable. \(\square \)

Lemma 37

Suppose that the filtration \(({\mathcal {F}}_t)_{t\in [0,T]}\) is right-continuous. Let \((M_t)_{t\in [0,T]}\) be a right-continuous sub-maxingale. Let \(\tau \), S be two stopping times such that \(\tau (\Omega )\) is a finite set. Then, we have \({{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{S}}(M_{\tau })\ge M_{\tau \wedge S}\).

Proof

Let \((S_n)_n\) be a sequence of stopping times decreasing to S as given in Lemma 36. Recall that \(S_n(\Omega )\) is finite for all n. Moreover, we have \( {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{s}}(M_{\tau }) =\lim _n \uparrow {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{s_n}}(M_{\tau })\). By Lemma 35, we deduce that \( {{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{S}}(M_{\tau }) \ge \lim \uparrow M_{\tau \wedge S_n}.\) As \((\tau \wedge S_n)_n\) decreases to \(\tau \wedge S\) and M is right-continuous, we conclude that \({{\,\mathrm{ess\;sup}\,}}_{{\mathcal {F}}_{s}}(M_{\tau }) \ge M_{\tau \wedge s}\). \(\square \)

Appendix B: Auxiliary results

Lemma 38

Suppose that, at time \(t\le T\), \({\mathcal {O}}_t\) is the pseudo-distance topology defined by (8) and \({\mathcal {V}}_{t,T}^c={\mathcal {V}}_{t,T}^c({\mathcal {O}}_t)\). If \({\mathcal {V}}_{t,T}\) is \({\mathcal {F}}_t\)-decomposable, then \({\mathcal {V}}_{t,T}^c\) is \({\mathcal {F}}_t\)-decomposable.

Proof

Consider \(V_{t,T}^{c,i}\in {\mathcal {V}}_{t,T}^c\), \(i=1,2\), and \(F_t\in {\mathcal {F}}_t\). By Proposition 32, \(V_{t,T}^{c,i}\le V_{t,T}^{n,i}+\alpha _t^{n,i}\) where \(V_{t,T}^{n,i}\in {\mathcal {V}}_{t,T}\) and \(\alpha _t^{n,i}\) converges to 0 in probability as \(n\rightarrow \infty \), for \(i=1,2\). We set

$$\begin{aligned} V_{t,T}^{n}= & {} V_{t,T}^{n,1}1_{F_t}+V_{t,T}^{n,2}1_{\Omega \setminus F_t}, \quad \alpha _t^{n}=\alpha _t^{n,1}1_{F_t}+\alpha _t^{n,2} 1_{\Omega \setminus F_t}. \end{aligned}$$

Note that \(V_{t,T}^{n}\in {\mathcal {V}}_{t,T}\) by assumption and \(\alpha _t^{n}\) converges to 0 in probability. Moreover, \(V_{t,T}^{c,1}1_{F_t}+V_{t,T}^{c,2}1_{\Omega {\setminus } F_t}\le V_{t,T}^{n}+\alpha _t^{n}\). Therefore, Proposition 32 implies that \(V_{t,T}^{c,1}1_{F_t}+V_{t,T}^{c,2}1_{\Omega {\setminus } F_t}\in {\mathcal {V}}_{t,T}^c\) and the conclusion follows. \(\square \)

Lemma 39

Let \(h_T\in L^0(\textbf{R},{\mathcal {F}}_T)\) be a payoff. If \({\mathcal {V}}_{t,T}\) (resp. \({\mathcal {V}}_{t,T}^c\)) is \({\mathcal {F}}_t\)-decomposable (resp. infinitely \({\mathcal {F}}_t\)-decomposable), then \({\mathcal {P}}_{t,T}(h_T)\) (resp. \({\mathcal {P}}_{t,T}^c(h_T)\)) is \({\mathcal {F}}_t\)-decomposable (resp. infinitely \({\mathcal {F}}_t\)-decomposable).

Proof

Suppose that \({\mathcal {V}}_{t,T}\) is \({\mathcal {F}}_t\)-decomposable and consider \(p_t^1,p_t^2\in {\mathcal {P}}_{t,T}(h_T)\) and \(F_t\in {\mathcal {F}}_t\). Then, \(p_t^i+V_{t,T}^{i}\ge h_T\) for some \(V_{t,T}^{i}\in {\mathcal {V}}_{t,T}\), \(i=1,2\). By assumption, we have \(V_{t,T}=V_{t,T}^{1}1_{F_t} +V_{t,T}^{2}1_{\Omega {\setminus } F_t}\in {\mathcal {V}}_{t,T}\) by assumption and \(p_t^11_{F_t}+p_t^21_{\Omega \setminus F_t}+V_{t,T}\ge h_T\). We deduce that \(p_t^11_{F_t}+p_t^2 1_{\Omega {\setminus } F_t}\in {\mathcal {P}}_{t,T}(h_T)\). By the same reasoning, the property holds for \({\mathcal {V}}_{t,T}^c\) and the infinite \({\mathcal {F}}_t\)-decomposability is obtained similarly. The conclusion follows. \(\square \)

Lemma 40

Let \(h_T\in L^0(\textbf{R},{\mathcal {F}}_T)\) be a payoff. If \({\mathcal {V}}_{t,T}\) is infinitely \({\mathcal {F}}_t\)-decomposable, then for any \(\gamma _t\in L^0(\textbf{R},{\mathcal {F}}_t)\) such that \(\gamma _t>\pi _{t,T}(h_T)\), there exists a price \(p_t\in {\mathcal {P}}_{t,T}(h_T)\) such that \(p_t<\gamma _t\). In particular, \(\gamma _t\in {\mathcal {P}}_{t,T}(h_T)\).

Proof

Since \({\mathcal {V}}_{t,T}\) is infinitely \({\mathcal {F}}_t\)-decomposable, \({\mathcal {P}}_{t,T}(h_T)\) is infinitely \({\mathcal {F}}_t\)-decomposable by Lemma 39. Therefore, \({\mathcal {P}}_{t,T}(h_T)\) is directed downward and we deduce that \(\pi _{t,T}(h_T)\lim _n\downarrow p_t^n\) where \(p_t^n\in {\mathcal {P}}_{t,T}(h_T)\), see Kabanov and Safarian (2009, Sect. 5.3.1). Then, a.s.(\(\omega \)), there exits \(n(\omega )\) such that \(p_t^n(\omega )<\gamma _t(\omega )\). We then define

$$\begin{aligned} N_t= & {} \inf \{n\ge 1:~p_t^n<\gamma _t\}\in L^0({\mathbb {N}},{\mathcal {F}}_t),\\ p_t= & {} \sum _{j=1}^\infty p_t^j1_{\{N_t=j\}}. \end{aligned}$$

By assumption \(p_t\in {\mathcal {P}}_{t,T}(h_T)\) and \(p_t<\gamma _t\). The conclusion follows. \(\square \)

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Cherif, D., Lépinette, E. No-arbitrage conditions and pricing from discrete-time to continuous-time strategies. Ann Finance 19, 141–168 (2023). https://doi.org/10.1007/s10436-023-00426-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10436-023-00426-1

Keywords

JEL Classification

Navigation