1 Introduction

The main feature of the Black–Scholes market is that the price of a share of stock is given by the exponential Brownian motion. Thus, in particular, the logarithm of the price is a process with independent increments. However, this assumption is quite unrealistic, since the previous history of the market strongly influences its future behaviour. Therefore, there were several attempts to describe a market whose log-price is a stochastic process with dependent increments.

From the point of view of the theory of stochastic processes, the fractional Brownian motion is one of the most studied stochastic processes with dependent increments [1,2,3]. This process is, in particular, a Gaussian process. So, it was natural to study a Black–Scholes market in which the standard Brownian motion is replaced by a fractional Brownian motion. However, it was proved by Shiryaev [4] that such a market admits arbitrage opportunities.

Hu and Øksendal [5] proposed a model of Black–Scholes type which involves fractional Brownian motion and in which usual multiplication is replaced by the so-called Wick product. They proved that such a market has no arbitrage opportunities, which is complete, and there is an analogue of the Black–Scholes pricing formula for a European call option. We also refer to [6] for an extension of the result of Hu and Øksendal.

In recent years, the fractional Brownian model has been widely used in finance, for example, many consider fractional Brownian motion model in option pricing [7,8,9,10,11,12,13,14]. But none of them balance the rigorousness of the theory and application of fractional Brownian motion well. The aim of this study is to make a systematic introduction into the fractional white noise calculus and to explain the main ideas of Hu and Øksendal’s result [5]. In particular, general Gaussian measures on the dual of a nuclear space are studied. Additionally to [5], some existing results from the book by Berezansky and Kondratiev [15] are taken. Moreover, correspondent proofs of the above literature are strongly expanded and the asset price movement is simulated based on fractional Brownian motion.

Our paper is organized as follows: in Sect. 2, some results of fractional white noise calculus are briefly reviewed. In Sect. 3, the fractional Brownian motion is used to model the financial market and a pricing formula for European call option is derived based on fractional Brownian motion. In Sect. 4, empirical analysis between fractional Brownian motion and Brownian motion will be compared using Monte Carlo simulation. In Sect. 5, the results above are concluded. And in “Appendix”, fractional stochastic integral of Itô type is defined using wick product from Gel’fand triple.

2 Review of Fractional White Noise Calculus

In this section, we will shortly review some results of fractional white noise calculus. Some known results and detailed account of the underlying theory can be found in “Appendix”. The following two definitions are taken from [16].

Definition 1

Let \(F:S'({\mathbb {R}})\rightarrow {\mathbb {R}}\) and let \(\gamma \in S'({\mathbb {R}})\). We say that F has a directional derivative in the direction \(\gamma \) if

$$\begin{aligned} D_{\gamma }F(x):=\lim _{\varepsilon \rightarrow 0}\frac{F(x+\varepsilon \gamma )-F(x)}{\varepsilon } \end{aligned}$$
(1)

exists in \(({\tilde{S}})'\).

Example 1.1 If \(F(x)=\langle x,f\rangle \) for some \(f\in S(\mathbb R)\) and \(\gamma \in L^2({\mathbb {R}})\subset S'({\mathbb {R}})\) then

$$\begin{aligned} D_{\gamma }F(x)&=\lim _{\varepsilon \rightarrow 0}\frac{1}{\varepsilon }[\langle x+\varepsilon \gamma ,f\rangle -\langle x,f\rangle ]\\&=\lim _{\varepsilon \rightarrow 0}\frac{1}{\varepsilon }[\langle \varepsilon \gamma ,f\rangle ]=\langle \gamma ,f\rangle =\int _{\mathbb R}f(t)\gamma (t)dt. \end{aligned}$$

Definition 2

We say that \(F:S'({\mathbb {R}})\rightarrow {\mathbb {R}}\) is differentiable if there exists a map \(\Psi :{\mathbb {R}}\rightarrow ({\tilde{S}})'\) such that

$$\begin{aligned} \Psi (t)\gamma (t)=\Psi (t,x)\gamma (t) \end{aligned}$$

is \(({\tilde{S}})'\)-integrable and

$$\begin{aligned} D_{\gamma }F(x)=\int _{{\mathbb {R}}}\Psi (t,x)\gamma (t)dt \end{aligned}$$

for all \(\gamma \in L^2({\mathbb {R}})\). In this case, we put

$$\begin{aligned} D_{t}F(x):=\Psi (t,x) \end{aligned}$$

and call it the Malliavin (or Hida) derivative of F at t.

Remark

In fact, in the above definition one assumes that the product \(\Psi (t,x)\gamma (t)\) is well defined.

Example 1.2 Let \(F(x)=\langle x,f\rangle \) with \(f\in S({\mathbb {R}})\). Then, by Example 7.1 F is differentiable and its stochastic gradient is

$$\begin{aligned} D_{t}F(x)=f(t)\quad for\ a.a. (t,x). \end{aligned}$$

Let \({\mathfrak {F}}^{(H)}_{t}\) be the \(\sigma \)-algebra generated by \(B_{H}(s)\), \(0\le s\le t\). Let \(\mathfrak L^{1,2}_{\phi }({\mathbb {R}})\) denote the completion of the set of all \({\mathfrak {F}}^{(H)}_{t}\)-adapted processes \(f(t)=f(t,x)\) such that

$$\begin{aligned} \Vert f\Vert ^2_{{\mathfrak {L}}^{1,2}_{\phi }({\mathbb {R}})}:=\mathbb E_{\mu _{\phi }}\left[ \int _{{\mathbb {R}}}\int _{\mathbb R}f(s)f(t)\phi (s,t)dsdt\right] +{\mathbb {E}}_{\mu _{\phi }}\left[ \left( \int _{\mathbb R}D^{\phi }_{s}f(s)ds\right) ^2\right] <\infty ,\nonumber \\ \end{aligned}$$
(2)

where \(D^{\phi }_{s}\) is the \(\phi \) derivative defined in [17]. Namely, if \(D_{s}F\) denotes the usual Malliavin derivative, then

$$\begin{aligned} D^{\phi }_{s}F=\int _{{\mathbb {R}}}\phi (s,t)D_{t}Fdt. \end{aligned}$$

Then, by [18], Theorem 3.7 we have the following fractional Itô isometry:

$$\begin{aligned} {\mathbb {E}}_{\mu _{\phi }}\left[ \left( \int _{\mathbb R}f(t,x)dB_{H}(t)\right) ^2\right] =\Vert f\Vert ^2_{\mathfrak L^{1,2}_{\phi }({\mathbb {R}})}. \end{aligned}$$
(3)

Moreover, we have

$$\begin{aligned} {\mathbb {E}}_{\mu _{\phi }}\left[ \int _{{\mathbb {R}}}f(t,x)dB_{H}(t)\right] =0\quad for\ f\in {\mathfrak {L}}^{1,2}_{\phi }({\mathbb {R}}). \end{aligned}$$
(4)

Example 1.3 Let \(f,g\in H_{\phi }\). Then

$$\begin{aligned}:\langle f,x\rangle :\diamond :\langle g,x\rangle :=:\langle f{\hat{\otimes }}g,x^{\otimes 2}\rangle :. \end{aligned}$$

By Theorem 9(i) in “Appendix” we have

$$\begin{aligned}:\langle f^{\otimes 2},x^{\otimes 2}\rangle :=\Vert f\Vert ^2_{H_{\phi }}H_2(\Vert h\Vert ^{-1}_{H_{\phi }}\langle h,x\rangle ) \end{aligned}$$

and

$$\begin{aligned} H_2(s)=s^2-1. \end{aligned}$$

Therefore, we obtain

$$\begin{aligned}:\langle f^{\otimes 2},x^{\otimes 2}\rangle :=\Vert f\Vert ^2_{H_{\phi }}(\Vert h\Vert ^{-1}_{H_{\phi }}\langle h,x\rangle )^2-1)=\langle h,x\rangle ^2-\Vert f\Vert ^2_{H_{\phi }}. \end{aligned}$$

Moreover, by the polarization identity, we have

$$\begin{aligned}:\langle f{\hat{\otimes }}g,x^{\otimes 2}\rangle :=\langle f,x\rangle \langle g,x\rangle -(f,g)_{H_{\phi }}, \end{aligned}$$

so that

$$\begin{aligned}:\langle f,x\rangle :\diamond :\langle g,x\rangle :=\langle f,x\rangle \langle g,x\rangle -(f,g)_{H_{\phi }}. \end{aligned}$$

Therefore, by Eqs. (5), (6), the standard Wick calculus and finally the fact that

$$\begin{aligned} \langle \chi _{[0,t]},\chi _{[0,t]}\rangle _{H_{\phi }}=\int ^{t}_{0}\int ^{t}_{0}\phi (u,v)dudv=t^{2H}, \end{aligned}$$

we have

$$\begin{aligned} \int ^{t}_{0}B_{H}(s)dB_{H}(s)&=\int ^{t}_{0}B_{H}(s)\diamond W_{H}(s)ds\nonumber \\&=\int ^{t}_{0}B_{H}(s)\diamond \frac{d}{ds}B_{H}(s)ds=\int ^{t}_0\frac{d}{ds}B^{\diamond 2}_{H}(s)ds\nonumber \\&=\frac{1}{2}B^{\diamond 2}_{H}(t)=\frac{1}{2}B^2_{H}(t)-\frac{1}{2}t^{2H}. \end{aligned}$$
(5)

Example 1.4 (Geometric fractional Brownian motion) Consider the fractional stochastic differential equation

$$\begin{aligned} dX(t)=\mu X(t)dt+\sigma X(t)dB_{H}(t),\quad X(0)=x>0, \end{aligned}$$
(6)

where x, \(\mu \) and \(\sigma \) are constants. By (158), we get:

$$\begin{aligned} dX(t)=\mu X(t)dt+\sigma X(t)\diamond W_{H}(t)dt. \end{aligned}$$

Dividing by dt, we obtain

$$\begin{aligned} \frac{dX(t)}{dt}=\mu X(t)+\sigma X(t)\diamond W_{H}(t). \end{aligned}$$

Since \(\mu \) is a constant

$$\begin{aligned} \mu X(t)=\mu \diamond X(t). \end{aligned}$$

Furthermore, since \(\diamond \) is commutative,

$$\begin{aligned} X(t)\diamond W_{H}(t)=W_{H}(t)\diamond X(t), \end{aligned}$$

and so

$$\begin{aligned} \frac{dX(t)}{dt}=(\mu +\sigma W_{H}(t))\diamond X(t). \end{aligned}$$

Using Wick calculus, we see that the solution of this equation is

$$\begin{aligned} X(t)=x\exp ^{\diamond }\left( \mu t+\sigma \int ^{t}_{0}W_{H}(s)ds\right) =x\exp ^{\diamond }(\mu t+\sigma B_{H}(t)). \end{aligned}$$
(7)

Here,

$$\begin{aligned} \exp ^{\diamond }F=1+\sum ^{\infty }_{n=1}\frac{1}{n!}\underbrace{F\diamond F\diamond \dots \diamond F}. \end{aligned}$$
(8)

Using (139) in “Appendix” and (5) we have

$$\begin{aligned} X(t)&=\exp ^{\diamond }(\mu t+\sigma B_{H}(t))\nonumber \\&=\exp ^{\diamond }(\mu t)\exp ^{\diamond }(\sigma B_{H}(t))\nonumber \\&=\exp (\mu t)\exp ^{\diamond }(\sigma B_{H}(t))\nonumber \\&=\exp (\mu t)\exp \left( \sigma B_{H}(t)-\frac{\sigma ^2}{2}|\chi _{[0,t]}|^2_{\phi }\right) \nonumber \\&=\exp (\mu t)\exp \left( \sigma B_{H}(t)-\frac{\sigma ^2}{2} t^{2H}\right) \nonumber \\&=x\exp \left( \sigma B_{H}(t)+\mu t-\frac{1}{2}\sigma ^2 t^{2H}\right) . \end{aligned}$$
(9)

The following results we have taken from the paper [5] without proofs.

Theorem 1

(Fractional Girsanov formula) Let \(T>0\) and let \(\gamma \) be a continuous function with \({\text {supp}}\gamma \subset [0,T]\). Let K be a function with \({\text {supp}}K\subset [0,T]\) and such that

$$\begin{aligned} (K,f)_{H_{\phi }}=(\gamma ,f)_{L^2({\mathbb {R}})},\quad \forall f\in S({\mathbb {R}}),\quad {\text {supp}} f\subset [0,T], \end{aligned}$$
(10)

i.e.

$$\begin{aligned} \int _{{\mathbb {R}}}K(s)\phi (s,t)ds=\gamma (t),\quad 0\le t\le T. \end{aligned}$$
(11)

Define a probability measure \(\mu _{\phi ,\gamma }\) on the \(\sigma \)-algebra \({\mathcal {F}}^{(H)}_{T}\) generated by\(\{B_{H}(s):0\le s\le T\}\) by

$$\begin{aligned} \frac{d\mu _{\phi ,\gamma }}{d\mu _{\phi }}=\exp ^{\diamond }\{-\langle x,K\rangle \}. \end{aligned}$$
(12)

Then, \({\hat{B}}_{H}(t)=B_{H}(t)+\int ^{t}_{0}\gamma _{s}ds\), \(0\le t\le T\), is a fractional Brownian motion under \(\mu _{\phi ,\gamma }\).

Remark 1

Recall that, in the classical case, the Girsanov theorem allows us to find a probability measure which is equivalent to the initial probability measure under which a "shifted" Brownian motion becomes a usual Brownian motion. In the fractional case, the above theorem gives a similar result, under a new probability measure \(\mu _{\phi , \gamma }\) (which is equivalent to the initial measure \(\mu _{\phi }\)) the shifted fractional Brownian motion \({\hat{B}}_{H}(t)=B_{H}(t)+\int ^{t}_{0}\gamma _{s}ds\) becomes a usual fractional Brownian motion.

Remark 2

Since \(B_{H}(t)\) is not a martingale, unlike in the standard Brownian motion case, the restriction of \(\frac{d\mu _{\phi ,\gamma }}{d\mu _{\phi }}\) to \({\mathcal {F}}^{(H)}_{t}\), \(0<t<T\), is in general not given by \(\exp ^{\diamond }\{-\langle \omega ,\chi _{[0,t]}K\rangle \}.\)

Proposition 1

(Wick products on different white noise spaces) Let \(P=\mu _{\phi }\), \(Q=\mu _{\phi ,\gamma }\) and \({\hat{B}}_{H}(t)=B_{H}(t)+\int ^{t}_{0}\gamma _{s}ds\) be as in Theorem 1. Let the Wick products corresponding to P and Q be denoted by \(\diamond _{p}\) and \(\diamond _{Q}\), respectively. Then

$$\begin{aligned} F\diamond _{p}G=F\diamond _{Q}G \end{aligned}$$

for all \(F,G\in ({\tilde{S}})'\).

Definition 3

Let \(G=:\langle g_{n}, \omega ^{\otimes n}\rangle :\), where \(g_n\in H^{\otimes n}_{\phi }\). Then, we define the fractional (or quasi-) conditional expectation of G with respect to \({\mathcal {F}}^{(H)}_{t}\) by

$$\begin{aligned} \tilde{\mathbb E}_{\mu _{\phi }}[G\mid {\mathcal {F}}^{(H)}_{t}]=:\langle g_n 1^{\otimes n}_{[0,t]},\omega ^{\otimes n}\rangle :, \end{aligned}$$
(13)

and we extend \(\tilde{\mathbb E}_{\mu _{\phi }}(\cdot |{\mathcal {F}}^{(H)}_{t})\) by linearity.

Remark

In the classical case, this would be indeed the conditional expectation with respect to \({\mathcal {F}}_t\). In the fractional case, this is not so, i.e. \(\tilde{{\mathbb {E}}}(\cdot |\mathcal F^{(n)}_{t})\ne {\mathbb {E}}(\cdot |{\mathcal {F}}^{(n)}_{t}).\)

Theorem 2

(A fractional Clark–Ocone theorem) Suppose that \(G(x)\in L^2(\mu _{\phi })\) is \({\mathcal {F}}^{(H)}_{T}\)-measurable. Define

$$\begin{aligned} \Psi (t,x):=\tilde{{\mathbb {E}}}[D_{t}G\mid {\mathcal {F}}^{(H)}_{t}]. \end{aligned}$$

Then \(\Psi \) belong to the space \(\mathcal L^{1,2}_{\phi }({\mathbb {R}})\) and we have

$$\begin{aligned} G(x)={\mathbb {E}}_{\mu _{\phi }}[G]+\int ^{T}_{0}\tilde{\mathbb E}_{\mu _{\phi }}[D_{t}G\mid {\mathcal {F}}^{(H)}_{t}]dB_{H}(t). \end{aligned}$$

We shall call \(\tilde{{\mathbb {E}}}_{\mu _{\phi }}[D_{t}G|\mathcal F^{(H)}_{t}]\) the fractional Clark derivative of G in analogy with the classical Brownian motion case. We will use the notation

$$\begin{aligned} \nabla ^{\phi }_{t}G={\tilde{E}}_{\mu _{\phi }}[D_{t}G\mid {\mathcal {F}}^{(H)}_{t}], \end{aligned}$$

so that

$$\begin{aligned} G(x)={\mathbb {E}}_{\mu _{\phi }}[G]+\int ^{T}_{0}\nabla ^{\phi }_{t}GdB_{H}(t). \end{aligned}$$

Remark 1 By analogy with the classical case, the fractional Clark–Ocone formula gives a representation of a square-integrable random variable G which is defined through the information available up to time T (i.e. \({\mathfrak {F}}^{(H)}_{T}\)-measureable) as the expectation of G plus the integral from 0 till T with respect to the fractional Brownian motion in which the integrand for each \(t\in [0,T]\) depends only on the information available up to time t (i.e. \(\mathfrak F^{(H)}_{t}\)-measurable).

3 Fractional Black–Scholes Market

We will now consider the Black–Scholes market with fractional Brownian motion. This market has a bank account or a bond, where the price A(t) at time t is described by the equation

$$\begin{aligned} dA(t)=\rho A(t)dt,\quad A(0)=1,\quad 0\le t\le T, \end{aligned}$$
(14)

where \(\rho >0\) is constant, i.e. \(A(t)=e^{\rho t}\) and a stock whose price X(t) at time t satisfies the equation

$$\begin{aligned} dX(t)=\mu X(t)dt+\sigma X(t)dB_{H}(t), \quad X(0)=\varkappa >0, \end{aligned}$$
(15)

where \(\mu \) and \(\sigma \ne 0\) are constants, \(0\le t\le T\). By Example 1.4 we have

$$\begin{aligned} X(t)=x\exp \left( \sigma B_{H}(t)+\mu t-\frac{1}{2}\sigma ^2 t^{2H}\right) ,\quad t\ge 0. \end{aligned}$$
(16)

A portfolio or trading strategy \(\theta (t)=\theta (t,x)=(u(t),v(t))\) is an \({\mathcal {F}}^{(H)}_{t}\)-adapted 2-dimensional process giving the number of units u(t), v(t) held at time t of the bond and the stock, respectively.

We assume that the corresponding value \(Z(t)=Z^{\theta }(t,\omega )\) is given by

$$\begin{aligned} Z^{\theta }(t,\omega )=u(t)A(t)+v(t)\diamond X(t). \end{aligned}$$
(17)

Remark

Note that in the above formula, one replaces the usual multiplication by the Wick product. This will later on allow no arbitrage opportunities. However, it is not completely clear whether the above definition is indeed appropriate from the point of view of finance. The same remains in force for some definitions discussed below,

The portfolio is called self-financing if

$$\begin{aligned} dZ^{\theta }(t,\omega )&=u(t)dA(t)+v(t)\diamond dX(t)\nonumber \\&=u(t)dA(t)+\mu v(t)\diamond X(t)dt+\sigma v(t)\diamond X(t)dB_{H}(t);\quad t\in [0,T]. \end{aligned}$$
(18)

We assume that \(\theta =(u,v)\) is a self-financing portfolio. Then, by (17) we have

$$\begin{aligned} u(t)=\frac{Z^{\theta }(t)-v(t)\diamond X(t)}{A(t)}. \end{aligned}$$
(19)

Substituting (14) and (146) in “Appendix” into Eq. (18) gives

$$\begin{aligned} dZ^{\theta }(t)=\rho Z^{\theta }(t)dt+\sigma v(t)\diamond X(t)[\frac{\mu -\rho }{\sigma }dt+dB_{H}(t)]. \end{aligned}$$
(20)

By the Girsanov theorem for fractional Brownian motion (Theorem 1) we see that

$$\begin{aligned} {\hat{B}}_{H}(t):=\frac{\mu -\rho }{\sigma }t+B_{H}(t),\quad t\in [0,T], \end{aligned}$$
(21)

is a fractional Brownian motion with respect to the measure \({\hat{\mu }}_{\phi }\) defined on \({\mathcal {F}}^{(H)}_{T}\) by

$$\begin{aligned} d{\hat{\mu }}_{\phi }(x)=\exp \left( -\int ^{T}_{0}K(s)dB_{H}(s)-\frac{1}{2}\Vert K\Vert ^2_{\phi }\right) d\mu _{\phi }(x), \end{aligned}$$
(22)

where \(K(s)=K(T,x)\) is defined by the following properties: \({\text {supp}} K\subset [0,T]\) and

$$\begin{aligned} \int ^{T}_{0}K(T,s)\phi (t,s)ds=\frac{\mu -\rho }{\sigma },\quad for\ 0\le t\le T. \end{aligned}$$
(23)

By [5], the solution of (23) is given by

$$\begin{aligned} K(T,s)=\frac{\mu -\rho }{2\sigma H(2H-1)\Gamma (2H-1)\Gamma (2-2H)\cos (\pi (H-\frac{1}{2}))}(Ts-s^2)^{\frac{1}{2}-H}, \nonumber \\ \end{aligned}$$
(24)

where \(\Gamma (\cdot )\) denotes the gamma function.

In terms of \({\hat{B}}_{H}(t)\), Eq. (20) can be written as

$$\begin{aligned} dZ^{\theta }(t)=\rho Z^{\theta }(t)dt+\sigma v(t)\diamond X(t)d{\hat{B}}_{H}(s),\quad 0\le t\le T, \end{aligned}$$
(25)

Multiplying by \(e^{-\rho t}\) and integrating, we get

$$\begin{aligned} e^{-\rho t}Z^{\theta }(t)=e^{-\rho t}Z^{\theta (t),z}=z+\int ^{t}_{0}e^{-\rho s}\sigma v(s)\diamond X(s)d{\hat{B}}_{H}(s),\quad 0\le t\le T,\nonumber \\ \end{aligned}$$
(26)

where \(z=Z^{\theta }(0)\)(constant) is the initial fortune.

Definition 4

We say that the portfolio is admissible if it is self-financing and

$$\begin{aligned} v\diamond X\in \hat{\mathfrak {L}}^{1,2}_{\phi }({\mathbb {R}}) \end{aligned}$$
(27)

where \(\hat{\mathfrak {L}}^{1,2}_{\phi }({\mathbb {R}})\) is the space defined similarly to \(\mathfrak {L}^{1,2}_{\phi }({\mathbb {R}})\)(see Eq. (2)), but with \(B_{H}(t)\), \(\mu _{\phi }\) replaced by \({\hat{B}}_{H}(t)\), \({\hat{\mu }}_{\phi }.\)

Definition 5

An admissible portfolio \(\theta \) is called an arbitrage if

$$\begin{aligned} Z^{\theta }(0)\le 0,\quad Z^{\theta }(T)\ge 0\quad a.s.\ and\quad \mu _{\phi }({x:Z^{\theta }(T,x)}>0)>0. \end{aligned}$$
(28)

From Eq. (26) we deduce that no arbitrage opportunities can exist. Indeed, by taking the expectation with respect to \({\hat{\mu }}_{\phi }\), we get by Proposition 1 and Eq. (4) applied to \({\hat{\mu }}_{\phi }\):

$$\begin{aligned} {\mathbb {E}}_{{\hat{\mu }}_{\phi }}[e^{-\rho T}Z^{\theta }(T)]=z+\mathbb E\left( \int ^{t}_{0}e^{-\rho s}\sigma v(s)\diamond X(s)d{\hat{B}}_{H}(s)\right) =z=Z^{\theta }(0). \end{aligned}$$
(29)

Thus, if \(Z^{\theta }(T)\ge 0\) and \(\mu _{\phi }(x: Z^{\theta }(T,x)>0)>0\), then since \({\hat{\mu }}_{\phi }\) is equivalent to \(\mu \), \({\mathbb {E}}_{{\hat{\mu }}_{\phi }}(e^{-\rho T}Z^{\theta }(T))>0\) and so \(Z^{\theta }(0)>0\).

Definition 6

The market (A(t), X(t)); \(t\in [0,T]\) is called complete if for every \({\mathfrak {F}}^{(H)}_{T}\)-measurable bounded random variable F(x) there exist \(z\in {\mathbb {R}}\) and portfolio \(\theta =(u,v)\) such that

$$\begin{aligned} F(x)=Z^{\theta ,z}(T,x)\quad \mu _{\theta }-a.s. \end{aligned}$$
(30)

By Eq. (26), (30) is equivalent to

$$\begin{aligned} e^{-\rho T}F(x)=z+\int ^{T}_{0}e^{-\rho t}\sigma v(t)\diamond X(t)d{\hat{B}}_{H}(t). \end{aligned}$$
(31)

If we apply the fractional Clark–Ocone theorem (Theorem 2) to \(G(x)=e^{-\rho T}F(x)\) and with \(B_{H}(t)\) replaced by \({\hat{B}}_{H}(t)\), we get

$$\begin{aligned} e^{-\rho T}F(x)={\mathbb {E}}_{{\hat{\mu }}_{\phi }}[e^{-\rho T}F]+\int ^{T}_{0}\tilde{{\mathbb {E}}}_{{\hat{\mu }}_{\phi }}[e^{-\rho T}\hat{D_{t}}|\mathfrak {F}^{(H)}_{t}]d{\hat{B}}_{H}(t), \end{aligned}$$
(32)

where \({\hat{D}}_{t}\) denotes the stochastic gradient with respect to \({\hat{\mu }}_{\phi }\). Note that the \(\sigma \)-algebra \(\hat{\mathfrak F}^{(H)}_{t}\) generated by \({\hat{B}}_{H}(s)\), \(s\le t\), is the same as \({\mathfrak {F}}^{(H)}_{t}.\)

Comparing (31) and (32), we conclude that our market is indeed complete. Indeed, choose

$$\begin{aligned} z=Z^{\theta }(0)={\mathbb {E}}_{{\hat{\mu }}_{\theta }}[e^{-\rho T}F] \end{aligned}$$

and find v from the equation

$$\begin{aligned} e^{-\rho T}\sigma v(t)\diamond X(t)=\tilde{{\mathbb {E}}}_{{\hat{\mu }}_{\phi }}(e^{-\rho T}F|\mathfrak F^{(H)}_t). \end{aligned}$$

Thus,

$$\begin{aligned} v(t)=e^{\rho T}\sigma ^{-1}\tilde{{\mathbb {E}}}_{\mu _{\phi }}(e^{-\rho T}F|{\mathfrak {F}}^{(H)}_{t})\diamond X(t)^{\diamond (-1)}, \end{aligned}$$

and finally choose u(t) according to (146) in “Appendix”. If F represents a pay-off of the contingent claim at time T, then the time 0 fair price of this claim must be the time 0 value of the trading strategy which replicates this claim, i.e. z. Thus, we have proved the following theorem:

Theorem 3

The fractional Black–Scholes market has no arbitrage opportunities. It is complete and the prize z of a \({\mathfrak {F}}^{(H)}_{T}\)-measurable claim \(F(x)\in L^2({\hat{\mu }}_{\phi })\) is given by

$$\begin{aligned} z=e^{-\rho T}{\mathbb {E}}_{{\hat{\mu }}_{\phi }}[F], \end{aligned}$$
(33)

where \(\mu _{\phi }\) is defined in (22). Moreover, the corresponding replicating portfolio \(\theta (t)=(u(t),v(t))\) for the claim F is

$$\begin{aligned} v(t)=e^{-\rho (T-t)}\sigma ^{-1}X^{\diamond (-1)}(t)\diamond \tilde{\mathbb E}_{{\hat{\mu }}_{\phi }}[{\hat{D}}_{t}F\mid {\mathfrak {F}}^{(H)}_t] \end{aligned}$$
(34)

and u(t) is determined by (146).

Let us now consider a European call option [19]:

$$\begin{aligned} F(x)=(X(T,x)-c)^{+}, \end{aligned}$$
(35)

where \(c>0\) is a constant (the exercise price). In this case we get:

Corollary 1

(The fractional Black–Scholes formula) The time 0 price of the fractional European call (35) is

$$\begin{aligned} z&=e^{-\rho T}\int _{{\mathbb {R}}}\frac{1}{T^{H}\sqrt{2\pi }}(x\exp [\sigma y+\rho T-\frac{1}{2}\sigma ^2 T^{2H}]-c)^{+}\exp [-\frac{y^2}{2T^{2H}}]dy \end{aligned}$$
(36)
$$\begin{aligned}&=x\varPhi \left( \eta +\frac{1}{2}\sigma T^{H}\right) -ce^{-\rho T}\varPhi \left( \eta -\frac{1}{2}\sigma T^{H}\right) , \end{aligned}$$
(37)

where

$$\begin{aligned} \eta =\sigma ^{-1}T^{-H}\left( {\text {log}}\frac{x}{c}+\rho T\right) \end{aligned}$$

and

$$\begin{aligned} \varPhi (t)=\frac{1}{\sqrt{2\pi }}\int ^{t}_{-\infty }e^{-\frac{s^2}{2}}ds \end{aligned}$$

is the normal distribution function (the error function).

The corresponding replication portfolio \(\theta (t)=(u(t),v(t))\) is given by Eq. (146) in “Appendix”, (26), and

$$\begin{aligned} v(t)=\sigma ^{-1}e^{-\rho (T-t)}X^{\diamond (-1)}(t)\diamond \Psi (X(t)), \end{aligned}$$
(38)

where

$$\begin{aligned} \Psi (y)=\frac{1}{\sqrt{2\phi (T^{2H}-t^{2H})}}\int _{\mathbb R}\exp \left( -\frac{({\tilde{y}}-z)^2}{\sqrt{2(T^{2H}-t^{2H})}}\right) h(z)dz. \end{aligned}$$
(39)

Here

$$\begin{aligned} {\tilde{y}}=\frac{\log y-\log x-\rho t+\frac{1}{2}\sigma ^2 t^{2H}}{\sigma }, \end{aligned}$$
(40)

where

$$\begin{aligned} h(z)=\sigma \chi _{[c,\infty ]}(x\exp \left( \sigma z+\rho T-\frac{1}{2}\sigma ^2 T^{2H}\right) . \end{aligned}$$
(41)

Proof

We will only prove the formulas (36) and (37), since the second part of the corollary uses one technical result which requires a long proof.

So, by (16), (21) and (33) we have

$$\begin{aligned} z&=e^{-\rho T}{\mathbb {E}}_{{\hat{\mu }}_{\phi }}[F]\nonumber \\&=e^{-\rho T}{\mathbb {E}}_{{\hat{\mu }}_{\phi }}[(X(T,x)-c)^{+}]\nonumber \\&=e^{-\rho T}{\mathbb {E}}_{{\hat{\mu }}_{\phi }}[(x\exp (\sigma B_{H}(T)+\mu T-\frac{1}{2}\sigma ^2 T^{2H})-c)^{+}]\nonumber \\&=e^{-\rho T}{\mathbb {E}}_{{\hat{\mu }}_{\phi }}[(x\exp (\sigma {\hat{B}}_{H}(T)+\rho T-\frac{1}{2}\sigma ^2 T^{2H})-c)^{+}]\nonumber \\&=e^{-\rho T}{\mathbb {E}}_{\mu _{\phi }}[(x\exp (\sigma B_{H}(T)+\rho T-\frac{1}{2}\sigma ^2 T^{2H})-c)^{+}]\nonumber \\&=e^{-\rho T}\int _{{\mathbb {R}}}\frac{1}{T^{H}\sqrt{2\pi }}(x\exp (\sigma y+\rho T-\frac{1}{2}\sigma ^2 T^{2H})-c)^{+}\nonumber \\&\times \exp [-\frac{y^2}{2T^{2H}}]dy \end{aligned}$$
(42)

which is (36).

Now, let us prove that (36) is equal to (37). Note that

$$\begin{aligned} x\exp [\sigma y+\rho T-\frac{1}{2}\sigma ^2 T^{2H}]-c\ge 0 \end{aligned}$$

if only if

$$\begin{aligned} \exp [\sigma y+\rho T-\frac{1}{2}\sigma ^2 T^{2H}]&\ge \frac{c}{x},\\ \sigma y+\rho T-\frac{1}{2}\sigma ^2 T^{2H}&\ge \log \frac{c}{x},\\ y&\ge \frac{\log \frac{c}{x}+\frac{1}{2}\sigma ^2 T^{2H}-\rho T}{\sigma }=:y_0. \end{aligned}$$

Then,

$$\begin{aligned} z&=e^{-\rho T}\int _{{\mathbb {R}}}\frac{1}{T^{H}\sqrt{2\pi }}\left( x\exp (\sigma y+\rho T-\frac{1}{2}\sigma ^2 T^{2H}\right) -c)^{+}\\&\quad \times \exp \left[ -\frac{y^2}{2T^{2H}}\right] dy\\&=e^{-\rho T}\int ^{\infty }_{y_0}\frac{1}{T^{H}\sqrt{2\pi }}\left( x\exp (\sigma y+\rho T-\frac{1}{2}\sigma ^2 T^{2H})-c\right) \\&\quad \times \exp \left[ -\frac{y^2}{2T^{2H}}\right] dy\\&=e^{-\rho T}\int ^{\infty }_{y_0}\frac{1}{T^{H}\sqrt{2\pi }}(x\exp (\sigma y+\rho T-\frac{1}{2}\sigma ^2 T^{2H}))\\&\quad \times \exp [-\frac{y^2}{2T^{2H}}]dy-c e^{-\rho T}\int ^{\infty }_{y_0}\frac{1}{T^{H}\sqrt{2\pi }}\exp \left[ -\frac{y^2}{2T^{2H}}\right] dy\\&=xe^{-\rho T}\int ^{\infty }_{y_0}\frac{1}{T^{H}\sqrt{2\pi }}\exp (-\frac{(y-\sigma T^{2H})^2}{2T^{2H}}+\rho T)dy\\&\quad -ce^{-\rho T}\int ^{\infty }_{\frac{y_0}{T^{H}}}\frac{1}{\sqrt{2\pi }}\exp [-\frac{y^2}{2}]dy\\&=xe^{-\rho T}\int ^{\infty }_{\frac{y_0-\sigma T^{2H}}{T^{H}}}\frac{1}{\sqrt{2\pi }}\exp (-\frac{y^2}{2}+\rho T)dy-ce^{-\rho T}\int ^{\infty }_{\frac{y_0}{T^{H}}}\frac{1}{\sqrt{2\pi }}\exp [-\frac{y^2}{2}]dy\\&=x\int ^{\infty }_{\frac{y_0-\sigma T^{2H}}{T^{H}}}\frac{1}{\sqrt{2\pi }}\exp (-\frac{y^2}{2})dy-ce^{-\rho T}\int ^{\infty }_{\frac{y_0}{T^{H}}}\frac{1}{\sqrt{2\pi }}\exp [-\frac{y^2}{2}]dy\\&=x\int ^{-\frac{y_0-\sigma T^{2H}}{T^{H}}}_{-\infty }\frac{1}{\sqrt{2\pi }}\exp (-\frac{y^2}{2})dy-ce^{-\rho T}\int ^{-\frac{y_0}{T^{H}}}_{-\infty }\frac{1}{\sqrt{2\pi }}\exp [-\frac{y^2}{2}]dy, \end{aligned}$$

where, in the last equality, we used that the function \(\frac{1}{2\pi }e^{-\frac{x^2}{2}}\) is symmetric with respect to the origin. Since \(-\frac{y_0-\sigma T^{2\,H}}{T^{H}}=\eta +\frac{1}{2}\sigma T^{H}\) and \(-\frac{y_0}{T^{H}}=\eta -\frac{1}{2}\sigma T^{H}\), we reach (37).

Remark 2 Note that z is independent of \(\mu \). One may compare this corollary with the standard results for the classical Brownian motion case. We see that we get the fractional Black–Scholes formula from the classical Black–Scholes formula by simply changing the volatility \(\sigma \) in the classical formula with \(\sigma T^{H-\frac{1}{2}}\).

4 Monte Carlo Simulation and Numerical Results

This section provides the parameter estimation and Monte Carlo simulation of the models under the fractional Geometric Brownian motion and Geometric Brownian motion.

4.1 Parameter Estimation

We consider the Hurst index H first. If \(H\in (0, \frac{1}{2})\), then the disjoint increments are positively correlated. The process shows the short term dependency. If \(H=\frac{1}{2}\), then the process is the classical Brownian motion. If \(H\in (\frac{1}{2}, 1)\), the disjoint increments are negatively correlated. The process shows the long-term dependency.

In this study, the R/S analysis (i.e. rescaled range analysis) is implemented to estimate the Hurst parameter, H. Given a time series \({Mi, 1\le i\le M}\) of asset prices. We first transform the time series into logarithmic return series of the length \(N=M-1\) through logarithm.

$$\begin{aligned} N_i ={\text {log}} (\frac{M_{i+1}}{M_{i}}), \, i=1,2,\cdots , M-1. \end{aligned}$$

Then, we separate the new series into A subsets of the equal length and the length of every subset is \(n=N/A\). The mean value of every subset \(e_a\), \(a=1,2,\ldots , A\) is

$$\begin{aligned} e_a =\frac{1}{n}\sum ^n_i M_{ai}, \end{aligned}$$

where \(M_{ai}\) are the elements of subset a. Next in every subset a, we calculate the total sum of the difference of the first k values \(( k=1,2,\ldots , n)\) corresponding to the mean value \(e_a\) of the subset:

$$\begin{aligned} X_{k,a}=\sum ^k_{i=1} (N_{i,a} -e_a), \, k=1,2,\ldots , n. \end{aligned}$$

Then, the range \(R_a\) of every subset a is

$$\begin{aligned} R_a =\max (X_{k,a} )-\min (X_{k,a} ), \, 1\le k\le n. \end{aligned}$$

And the standard deviation \(S_a\) is

$$\begin{aligned} S_a = \sqrt{\frac{1}{n}\sum ^n_{i=1} (M_{ai} - e_a)^2}, \, a=1,2, \ldots , A. \end{aligned}$$

Then, in every subset a, we calculate \(\frac{R_a}{S_a}\) and calculate the mean of the rescaled range \(\frac{R_a}{S_a}\) of A subsets with time increment n:

$$\begin{aligned} (R/S)_{n}=\frac{1}{A}\sum ^{A}_{a=1} \frac{R_a}{S_a}. \end{aligned}$$

As we increase the value of n, we get the logarithmic asset value series of rescaled range \((R/S)_n\). According to the definition of Hurst index H, \((R/S)_n\) is in proportion to \(n^H\), i.e.

$$\begin{aligned} (R/S)_n = C\times n^H, \end{aligned}$$
(43)

where C is the coefficient. Taking the logarithm on both sides of (43), we have

$$\begin{aligned} \ln (R/S)_n = \ln C+H\ln n. \end{aligned}$$
(44)

According to (44), the value of H can be calculated through regression analysis.

We download the S &P 500 data from Yahoo Finance. In this paper, we select the closing index of the 251 trading days for one year from October 31, 2019 to October 30, 2020. Microsoft Excel software is used to do the regression analysis. We get \(H=0.6002\). And the goodness of fit is \(R^2=0.9591\), which indicates that the fitting effect is good. Therefore, it is reasonable to use the fractional Brownian motion to simulate the movement of the stock index. In the following, in the simulation process, we take the value of H as 0.6, which is rounded to the nearest integer (Fig. 1).

Fig. 1
figure 1

Rescales range analysis of the S &P 500 from 31 October 2019 to 30 October 2020

We consider again Eq. (15) satisfied by the stock price process \(\lbrace X_t, t\ge 0\rbrace \):

$$\begin{aligned} dX_t=\mu X_t dt+\sigma X_t dB_H (t), \end{aligned}$$

where \(\mu \) and \(\sigma \) are the expected return of stocks and the volatility of stock prices, respectively. \(B_H (t)\) in Eq. (15) represents the fractional Brownian motion, which is further expressed as

$$\begin{aligned} dB_H (t)=\epsilon \sqrt{2Ht^{2H-1}dt}, \end{aligned}$$
(45)

where \(\epsilon \sim N(0,1)\). If we set \(G_t=\ln X_t\), then by Eqs. (15), (45), and (16), we get

$$\begin{aligned} dG_t&=\frac{\partial G_t}{\partial t}dt+\frac{\partial G_t}{\partial X_t}dX_t +\frac{1}{2}\frac{\partial ^2 G_t}{\partial X^2_t}(dX_t)^2\\&=\frac{1}{X_t} (\mu X_t dt+\sigma X_t dB_H (t))-\frac{1}{2X_t^2}\sigma _t^2 X_t^2 d(t^{2H})\\&=\mu dt+\sigma _t dB_H (t)-\frac{1}{2}\sigma _t^2 2Ht^{2H-1}dt\\&=(\mu -\sigma _t^2 Ht^{2H-1} )dt+\sigma _t\epsilon \sqrt{2Ht^{2H-1}dt}. \end{aligned}$$

By discretization, we have

$$\begin{aligned} \Delta G_t&=\ln X_{t+\Delta t}-\ln X_t=\ln \frac{X_{t+\Delta t}}{X_t}\\&=(\mu -\sigma _t^2 Ht^{2H-1} )\Delta t+\sigma _t\epsilon \sqrt{2Ht^{2H-1}\Delta t}. \end{aligned}$$

It is trivial to see that

$$\begin{aligned} \frac{\ln \frac{X_{t+\Delta t}}{X_t}-(\mu -\sigma _t^2 Ht^{2H-1} )\Delta t}{\sqrt{\sigma _t^2 2Ht^{2H-1}\Delta t}}\sim N(0,1), \end{aligned}$$

and

$$\begin{aligned}{} & {} {\text {E}}(\ln \frac{X_{t+\Delta t}}{X_t})=(\mu -\sigma _t^2 Ht^{2H-1} )\Delta t,\\{} & {} {\text {Var}}(\ln \frac{X_{t+\Delta t}}{X_t})=\sigma _t^2 2Ht^{2H-1}\Delta t. \end{aligned}$$

Solving the above equations, we have

$$\begin{aligned} \mu= & {} \frac{{\text {E}}(\ln \frac{X_{t+\Delta t}}{X_t})}{\Delta t}+\frac{1}{2}\frac{{\text {Var}}(\ln \frac{X_{t+\Delta t}}{X_t})}{\Delta t}, \end{aligned}$$
(46)
$$\begin{aligned} \sigma _t^2= & {} \frac{{\text {Var}}(\ln \frac{X_{t+\Delta t}}{X_t})}{2Ht^{2H-1}\Delta t}. \end{aligned}$$
(47)

Equations (46) and (47) will be used in the following simulation process to estimate the mean value \(\mu \) and the standard deviation \(\sigma \) of the stock logarithmic return.

4.2 Numerical Results

In the following, we simulate the trend of the stock prices. The following are the concrete steps:

Step 1 Using the closing index of the 251 trading days from October 31, 2019 to October 30, 2020, and the closing index of October 30, 2019, which is rounded to 3048, we get the estimates of the mean value \(\mu \) and the standard deviation \(\sigma \) of the S &P 500 index logarithmic return by Eqs. (46) and (47).

Step 2 Discretizing the S &P 500 price Index, we have

$$\begin{aligned} \Delta X_{t_i}=X_{t_i+\Delta t} - X_{t_i}=\mu X_{t_i} \Delta t +\epsilon _{t_i +\Delta t} \sigma _{t_i+\Delta t} X_{t_i} \sqrt{2H t_i^{2H-1}\Delta t}. \end{aligned}$$

Let \(t_i =i\Delta t\) and \(\Delta t=1\), it becomes

$$\begin{aligned} X_{i} -X_{i-1} =\mu X_{i-1}+\epsilon _{i}\sigma _{i} X_{i-1}\sqrt{2H(i-1)^{2H-1}}, \end{aligned}$$

where \(\epsilon _i\) and \(\sigma _i\) indicate the standard normal random variable and corresponding standard deviation generated the ith time. Namely, when \(i=1\), it is

$$\begin{aligned} X_1=(\mu +1)X_0. \end{aligned}$$

When \(i=2\), it is

$$\begin{aligned} X_2 = (\mu +1)X_1 +\epsilon _2 \sigma _2 X_1 \sqrt{2H}, \end{aligned}$$

where \(\epsilon _2\) and \(\sigma _2\) indicate the standard normal random variable and the second standard deviation generated the second time. When \(i=3\), it is

$$\begin{aligned} X_3 = (\mu +1)X_2 +\epsilon _3 \sigma _3 X_2 \sqrt{2H 2^{2H-1}}, \end{aligned}$$

where \(\epsilon _3\) and \(\sigma _3\) indicate the standard normal random variable and the third standard deviation generated the third time. Likewise, we can obtain the 251 simulation values.

Step 3 Using the Microsoft Excel, we simulate the movement of S &P 500 index. Then, we compare the real movement with simulated movement to determine whether the price of S &P 500 index obeys the fractional Brownian motion. The results are shown in Figs. 23, and 4. In the graphs, the blue track represents the real movement, and the red track represents the corresponding fractional Brownian motion simulated movement. It is easy to observe the simulation effect is quite ideal.

Fig. 2
figure 2

First simulation of the S &P 500 from 31 October 2019 to 30 October 2020

Fig. 3
figure 3

Second simulation of the S &P 500 from 31 October 2019 to 30 October 2020

Fig. 4
figure 4

Third simulation of the S &P 500 from 31 October 2019 to 30 October 2020

Furthermore, we need a measurement to evaluate the results of our simulation. Here, we use root mean square error (RMSE). The root mean square error is defined as

$$\begin{aligned} R=\sqrt{\frac{1}{N}\sum ^n_{i=1} (Y_i -{\hat{Y}}_i)^2}. \end{aligned}$$

where N is the number of the samples. In our Monte Carlo simulations, \(N=251\). \(Y_i\) and \({\hat{Y}}_i\) are the actual price and the Monte Carlo simulated price of the S &P 500 index, respectively. Therefore, R value corresponding to the simulated movement in Figs. 23, and 4 are 225.252, 220.420, and 243.65. It can be seen that the second simulation is the highest and the third simulation is the lowest. In addition, in order to compare the fitting goodness between the fractional Brownian motion method and the Brownian motion simulation method, we use the previous data, and we get that Brownian motion simulation’s R value is 282.408. It is concluded that the proposed method is superior to the ordinary standard Brownian motion method. Therefore, it verifies the theory we construct in the previous chapters.

5 Conclusion

In this paper, we systematically study the fractional Brownian motion theory based on Gel’fand triples and apply the fractional Brownian motion to financial markets. During the simulation of S& P 500 index, unknown parameters need to be estimated and the goodness of fit of the models are compared between fractional Brownian motion and traditional Brownian motion. The results show that fractional Brownian motion model is more accurate than Brownian motion model.

Future work can consider simulate the fractional Gaussian noise using more factors like mean-reverting jump-type model and seasonality model. Moreover, data-driven methods can also be used to replicate the financial markets more closely.