Research article Special Issues

Quantum option pricing and data analysis

  • The paper proposes to treat financial models using techniques of quantum mechanics. The methodology relies on the Dirac matrix formalism and the Feynman path integral approach. This leads us to reexamine in this framework the classical option pricing models of Cox-Ross-Rubinstein and Black-Scholes. Moreover, financial data are classified with respect to the spectrum of a certain observable and then analyzed to identify price jumps using supervised machine learning tools.

    Citation: Wenyan Hao, Claude Lefèvre, Muhsin Tamturk, Sergey Utev. Quantum option pricing and data analysis[J]. Quantitative Finance and Economics, 2019, 3(3): 490-507. doi: 10.3934/QFE.2019.3.490

    Related Papers:

    [1] Amir Ahmad Dar, N. Anuradha . Comparison: Binomial model and Black Scholes model. Quantitative Finance and Economics, 2018, 2(1): 230-245. doi: 10.3934/QFE.2018.1.230
    [2] Lichao Tao, Yuefu Lai, Yanting Ji, Xiangxing Tao . Asian option pricing under sub-fractional vasicek model. Quantitative Finance and Economics, 2023, 7(3): 403-419. doi: 10.3934/QFE.2023020
    [3] Yunjae Nam, Changwoo Yoo, Hyundong Kim, Jaewon Hong, Minjoon Bang, Junseok Kim . Accurate computation of Greeks for equity-linked security (ELS) near early redemption dates. Quantitative Finance and Economics, 2025, 9(2): 300-316. doi: 10.3934/QFE.2025010
    [4] Sebastian Ferrando, Andrew Fleck, Alfredo Gonzalez, Alexey Rubtsov . Trajectorial asset models with operational assumptions. Quantitative Finance and Economics, 2019, 3(4): 661-708. doi: 10.3934/QFE.2019.4.661
    [5] David Anderson, Urban Ulrych . Accelerated American option pricing with deep neural networks. Quantitative Finance and Economics, 2023, 7(2): 207-228. doi: 10.3934/QFE.2023011
    [6] Haoyu Wang, Dejun Xie . Optimal profit-making strategies in stock market with algorithmic trading. Quantitative Finance and Economics, 2024, 8(3): 546-572. doi: 10.3934/QFE.2024021
    [7] Manuela Larguinho, José Carlos Dias, Carlos A. Braumann . Pricing and hedging bond options and sinking-fund bonds under the CIR model. Quantitative Finance and Economics, 2022, 6(1): 1-34. doi: 10.3934/QFE.2022001
    [8] David Alaminos, M. Belén Salas, Ángela M. Callejón-Gil . Managing extreme cryptocurrency volatility in algorithmic trading: EGARCH via genetic algorithms and neural networks. Quantitative Finance and Economics, 2024, 8(1): 153-209. doi: 10.3934/QFE.2024007
    [9] Takashi Kanamura . Supply-side perspective for carbon pricing. Quantitative Finance and Economics, 2019, 3(1): 109-123. doi: 10.3934/QFE.2019.1.109
    [10] Per B. Solibakke . Forecasting hourly WTI oil front monthly price volatility densities. Quantitative Finance and Economics, 2024, 8(3): 466-501. doi: 10.3934/QFE.2024018
  • The paper proposes to treat financial models using techniques of quantum mechanics. The methodology relies on the Dirac matrix formalism and the Feynman path integral approach. This leads us to reexamine in this framework the classical option pricing models of Cox-Ross-Rubinstein and Black-Scholes. Moreover, financial data are classified with respect to the spectrum of a certain observable and then analyzed to identify price jumps using supervised machine learning tools.


    As pointed out in Claessens et al. (2014), global financial crises underline the importance of innovative modelling approaches to financial markets. The quantum mechanics approach suggests an alternative way to describe the unpredictable stock market behaviour (see e.g., Baaquie (2009)).

    This paper is motivated by the quantum mechanics approach to the actuarial modelling and risk analysis as initiated in Tamturk and Utev (2018) and developed in Lefèvre et al. (2018), Tamturk and Utev (2019). We modify and adapt their methods to derive and analyze a quantum-type financial modelling. Specifically, the Schrödinger, Heisenberg, Feynman, Dirac approaches in quantum mechanics (see, e.g., Griffiths and Schroeter (2018); Parthasarathy (2012); Plenio (2002)) will be applied to the well-known option pricing models of Black and Scholes (1973) and Cox et al. (1979).

    Moreover, motivated by the data analysis of the quantum reserve process proposed in Lefèvre et al. (2018), we model the financial data as eigenvalues of certain 1 or 2-step observable operators. These data then are analyzed to identify price jumps using supervised machine learning tools such as k-fold cross-validation techniques (see, e.g., Bishop (2006); Hastie et al. (2009); Wittek (2014)).

    Option pricing models using quantum techniques discussed, for example, in Baaquie(2004, 2014); Bouchaud and Potters (2003); Haven (2002) are often based on the Schrödinger wave function with Hamiltonian operator H and are mainly oriented to continuous-time markets. For discrete-time markets as considered here, following Chen(2001, 2004), , we choose the discrete-time formalism and analyze the quantum version of the Cox-Ross-Rubinstein binomial model. Then, we establish the limit of the spectral measures providing the convergence to the geometric Brownian motion model. We also identify the limit of the N-step non-self adjoint bond market as a planar Brownian motion.

    The paper is organized as follows. Section 2 deals with heterogeneous quantum binomial markets. In Section 3, two convergence properties to continuous-time quantum markets are obtained. In Section 4, discrete and continuous-time quantum mechanics techniques are applied to the problem of option pricing. Section 5 is devoted to an analysis of stock market data.

    The motivation to our models come from the quantum mechanics approach to insurance proposed in Tamturk and Utev (2018, 2019), the insurance claim data analysis via quantum tools introduced in Lefèvre et al. (2018) and the non-traditional financial modelling initiated in Ma and Utev (2012) and developed in Karadeniz and Utev(2015, 2018). The quantum modelling approach is partly inspired by Baaquie (2004); Chen(2001, 2004), and Parthasarathy (2012).

    We begin by outlining some standard arguments for a quantum-type modelling (see e.g., Lefèvre et al. (2018); Parthasarathy (2012)). An observable is a linear operator (matrix) on a certain Hilbert space. The quantum product of two independent observables A,B is implemented by the tensor product of the observables AB. So, ln(AB) acts as a quantum sum of two independent observables; in particular, AA is the quantum product of two independent identical observables.

    The 1-step quantum geometric random walk is defined as a 2×2 matrix A with eigenvalues eu,ed. Thus, AN (N0) models the N-step geometric random walk.

    The quantum type modification of the classical Cox-Ross-Rubinstein model is originated in Chen(2001, 2004). The dynamics per period is defined by two moves: eu (the share price goes up) and ed (the share price goes down) with d<0<u.

    The quantum binomial model over N periods is then represented by the share price operator

    HSN=S0HN, (1)

    where the main 1-step observable H=A is a 2×2 self-adjoint (hermitian) matrix with eigenvalues eu,ed and representation

    A=UDU=U(eu00ed)U, (2)

    where U is a 2×2 unitary matrix.

    In the sequel, the quantum binomial model discussed will be heterogeneous with share price operator

    HSN=S0H1HN. (3)

    The motivation to this relatively new financial model comes from Lefèvre et al. (2018) quantum mechanics approach to non-life insurance and the Lamplighter group approximation to the financial modelling suggested in Ma and Utev (2012). Based on the approach to the financial data analysis developed in Ma and Utev (2012); Karadeniz and Utev(2015, 2018) (see also references therein), we treat the data as having big jumps, say eu, small jumps, say ed or no jump. Moreover, the financial stock is observed at fixed times Δk but the number of jumps occurred during the time period ((k1)Δ,kΔ] is not observed. For simplicity and following Lefèvre et al. (2018), we then assume there are at most two jumps per period.

    In this circumstance, the main observable operator is given, similarly to the insurance case Lefèvre et al. (2018), by

    H3=S0(P0I4+P1(AI2)+P2(A2)), (4)

    where In is a n×n identity matrix, A is the 2×2 matrix representing the 1-step geometric move operator (2) with eigenvalues eu,ed, and P0,P1,P2 are 3×3 matrix projection operators corresponding to the 0,1,2 claim occurrence operators and defined by

    Pi=WDi+1|3W,i=0,1,2, (5)

    where Di+1|3 is a 3×3 diagonal matrix which has all its elements equal to 0 except the (i+1,i+1)-th with value 1, and W is a 3×3 unitary matrix.

    The share price operator over N periods is then defined as

    HSN=S0HN3=S0(P0I4+P1(AI2)+P2(A2))N. (6)

    Furthermore, when constructing the density operator, we consider the following two cases.

    Maxwell-Boltzmann statistics. In this case, the jump sizes are i.i.d. with a two-point distribution. More precisely, there are 0,1,2 jumps with probabilities δ0,δ1,δ2 given via a Poisson process, and each jump size has two possible values ed,eu with probabilities q,p (see also later in Section 4.1).

    Bose-Einstein statistics. In this case, the claim sizes are dependent but remain independent of the claim occurrences.

    This model makes a bridge between the traditional binomial model and the actuarial type model. In this case, the dynamics per period is defined by three moves: no change, down and up. The trinomial type financial modelling is a well established topic in finance (see e.g., Boyle (1986); Tian (1993); Leisen and Reimer (1996)).

    Now, the 1-step observable operator H2=B is the 3×3 self-adjoint matrix with eigenvalues eu,1,ed corresponding to these three moves and is given by

    H2=B=UDU=U(eu0001000ed)U, (7)

    where U is now a 3×3 unitary matrix.

    Then, the share price operator over N periods is defined by

    HSN=S0HN2=S0(U)NDNUN. (8)

    Remarks. Notice that the heterogeneous versions of the trinomial model and the actuarial type model are available with the representation similar to (3). However, they are not treated in this paper because the main purpose for considering these two models is the non-standard data analysis (to be presented in Section 5). Although the data analysis of time dependent models is a fascinating topic, it is out of the scope of this paper.

    As mentioned in above, the quantum binomial model (1) is originated in Chen(2001, 2004). Our presentation is somewhat different and based on the algebraic tensor product properties. In addition, since heterogeneity is an important topic in option pricing (see e.g., Benninga and Mayshar (2000)), we treat a slightly more general time dependent or heterogeneous quantum binomial model (3).

    A quantum state ρ is defined as a positive self-adjoint operator with trace tr(ρ)=1. We recall the following properties.

    Lemma 2.1. Let A,B,C,D be self-adjoint operators, U a unitary matrix, f a function of observables and ρ a quantum state. Then,

    tr(ABC)=tr(CAB),tr(AB)=tr(A)tr(B),(AB)(CD)=(AC)(BD),(AB)=AB,A,B0AB0,A=UDUf(A)=Uf(D)U,UρU=quantum state. (9)

    For simplicity, we choose a quantum state ρ as a tensor product, i.e.

    ρ=ρ1ρN, (10)

    where each ρi is a self-adjoint non-negative 2×2 matrix such that tr(ρi)=1. From Lemma 2.1, we see that ρ is a proper quantum state.

    The risk-neutral world of the quantum Black-Scholes model consists of self-adjoint non-negative 2×2 matrices ρi that satisfy

    tr(ρiHi)=1+ri,i=1,,N, (11)

    where ri is the risk-free interest rate for the period i.

    Let us define

    Hi=UiDiUi,and let~ρi=UiρiUi,i=1,,N. (12)

    Note that the ~ρi are also quantum states. In addition, ˜ρi has non-negative diagonal elements q(i)u,q(i)d and it can be shown to have the representation

    ˜ρi=(quixˉxqdi), (13)

    where

    qui=1+ridiuidi,qdi=1qui. (14)

    The transformed operator ˜ρ is then obtained from Lemma 2.1 (third property) as

    ˜ρ=U(ρ1...ρN)U=(U1UN)(ρ1...ρN)(U1UN)=(U1ρ1U1)(UNρNUN)=˜ρ1˜ρN, (15)

    after using (12), and ˜ρ is again a quantum state. For the classical probability case, that is when all matrices are commutative, this transform is the form of the change of measure.

    Moreover, the quantum no arbitrage condition (11) is satisfied for the transformed density, i.e.

    tr(~ρiDi)=1+ri,i=1,,N. (16)

    Consider a payoff function f and a discount factor

    dN=1/(1+r1)(1+rN). (17)

    From (3), the price of the general option without arbitrage OP(f(HSN)) for the N-step quantum binomial model is then defined by

    OP(f(HSN))=dNtr(ρf(HSN))=dNtr(ρ1...ρNf(S0H1...HN)). (18)

    Applying properties given in Lemma 2.1, we then obtain

    OP(f(HSN))=tr(ρ1...ρNf(S0H1...HN))(1+r1)(1+rN)=tr(ρ1...ρNUf(S0D1...DN)U)(1+r1)(1+rN)=tr([U(ρ1...ρN)U]f(S0D1...DN))(1+r1)(1+rN)=tr(˜ρ1˜ρNf(S0D1...DN))(1+r1)(1+rN), (19)

    after using (15). Since ˜ρ is a tensor product and f(S0D1...DN) is a diagonal matrix, we deduce the option pricing formula (20) below.

    Theorem 2.2. For the heterogeneous quantum binomial model,

    OP(f(HSN))=σf(S0yσ)qσ(1+r1)(1+rN), (20)

    where the index σ denotes any feasible path yσ of the form

    yσ=eσ1eσNwithσi{ui,di}, (21)

    which occurs with the probability qσ=qσ1qσN where qσi{qui,qdi}.

    In particular, for the homogeneous case (the quantum Cox-Ross-Rubinstein model) where for all i, ri=r,ui=u,di=d with qui=qu,qdi=qd, the formula (20) reduces to

    OP(f(HSN))=1(1+r)NNn=0f(S0euned(Nn))qndqNnu. (22)

    Non-self adjoint quantum binomial market. Assume that Hi are invertible 2×2 matrices with two different eigenvalues ed and eu but no more self-adjoint, in general. In this case the transformed density ˜ρi are no more proper states, in general. However, they still have same diagonal elements qui and qdi and so the transformed matrix ˜ρ defined in (15) again has same diagonal elements as in the self-adjoint case, but is not a proper state, in general. By inspecting the proof, we see that Theorem 2.2 holds true in this case as well.

    In this section, we consider the homogeneous quantum binomial model, and we discuss two examples on the convergence to continuous-time markets, namely the Black-Scholes model and the planar Brownian motion.

    Theorem 3.1. Let μN be the measure of the eigenvalues of HN with respect to the quantum state ρN. Suppose that r=λ/N and u=d=σN1/2. Then, as N,

    f(S0x)dμN(x)f(S0x)dμ(x), (23)

    where μ is a lognormal distribution (i.e. μ(x)=P(ea+σZx) for suitable constants a,σ and Z a standard normal variable).

    Proof. We begin with the representation via the spectral measure μN. Observe that

    Ef(HSN)=tr(ρNf(S0HN))=f(S0x)μN(dx), (24)

    where μN is the measure of the eigenvalues λσ of HN with respect to the quantum probability tr(ρNHN). Since λσ=eσ1eσN with σi{ui,di}, we get

    μN(x)=σ:λσxqσ=σ:eσ1eσNxqσ1qσN, (25)

    where qσi{qu,qd} respectively.

    Let us move on to the weak convergence desired. We can write that

    qσ1qσN=P(Y1=eσ1)P(YN=eσN)=P(Y1=eσ1,,YN=eσN), (26)

    where the Yi are i.i.d. variables with P(Yi=eu)=qu,P(Yi=ed)=qd. Thus,

    μN(x)=P(Y1YNx)=P(lnY1++lnYNlnx)=P(TNlnx), (27)

    with TNlnY1++lnYN. As N, we obtain from the central limit theorem that

    μN(x)=P(TNlnx)P(ea+σZx)=μ(x), (28)

    for certain constants a,b and Z a standard normal. This gives the limit result (23).

    Remarks. (ⅰ) The limiting measure μ corresponds to that of the geometric Brownian motion St=S0e(ρσ2/2)t+σBt when S0=1, t=1 and ρσ2/2=a.

    (ⅱ) It would be interesting to compare the technique with the semi-classical approximation, such as expanding the action around the classical path (see (48) in Contreras et al. (2010)). Another interesting question is to analyse connection with arbitrage as discussed in Haven (2002) and Contreras et al. (2010).

    (ⅲ) The Cauchy transform is an alternative approach to deal with quantum probabilities (see Mudakkar and Utev (2013)). In particular,the convergence of spectral measures is reduced to the convergence of the Cauchy transforms. We recall that the Cauchy transform for the measure μ is defined by

    Sμ(z)=Rμ(dt)tz,whereμ((a,b))=limϵ01πbaSμ(x+iϵ)dx, (29)

    for all zCR={zC:z0} and open intervals (a,b) with μ({a,b})=0. The goal is then to show that Sμn(z)Sμ(z). However,the common approach of the moment expansion does not work in this case since

    dμ(x)zx=k=01zk+1xkdμ(x)=k=01zk+1E(ek(a+σZ))=k=01zk+1eka+k2σ2/2=. (30)

    In this part,by treating the relatively simple bond market,we show that non-commutative markets provide a richer class of financial models (see also Haven (2002); Contreras et al. (2010); Herscovich (2016)).

    Bond market. Assume that returns are non-risky,that is the outcomes are equal. In the quantum setup,we suppose that matrix H has a single eigenvalue,eu say.

    Self-adjoint quantum bond market. Let ρN denote a quantum state defined as before. Assume that in addition operator H is self-adjoint. Then H=euI2 and

    HSN=S0HN=S0euNIN2andf(HSN)=f(S0euN)IN2. (31)

    In particular,any option claim f(HSN) is commutative with the state ρN which implies that the self-adjoint quantum bond market is commutative,that is equivalent to the simple classical probability financial market with two non-risky assets. The no-arbitrage condition becomes global to give eu=tr(ρH)=1+r since tr(ρ)=1 which restates that under no-arbitrage the non-risky returns are equal. Moreover,the option price for f(HSN) is provided by

    OP(f(HSN))=tr(ρNf(S0HN))(1+r)N=euNf(S0euN). (32)

    Non self-adjoint quantum bond market. Now,assume that H is no longer self-adjoint,for example a product of two noncommutative self adjoint observables. Thus,non self-adjoint quantum bond market is noncommutative,in general. In this case,we show that the limit is sensitive to the density state ρ and the representation of H. For simplicity,we assume that the basic observable H and the state ρ are defined by

    Hu=(eu100eu100eu),ρ=(ρ11yδyρ220δ0ρ33), (33)

    i.e.,Hu is now a 3×3 Jordan matrix. The bond price process is then defined by

    HSN=S0HNu. (34)

    To satisfy the no-arbitrage condition,we ask that

    tr(ρHu)=1+r,thusy=1+reu. (35)

    Theorem 3.2. Suppose that r=λ/N,u=a/N and δ=Δ/N with Δ0. Then,for any positive integer k,as N,

    OP([HSN]k)eλSk0E[ek(λ+Δ/2+iBΔ)], (36)

    regardless of a, where i is the imaginary unit and BΔ is a Brownian motion at time Δ.

    Proof. Notice that

    Hku=(ekue(k1)uke(k2)uk(k1)/20ekue(k1)uk00eku). (37)

    From the definition of HSN and ρ, we then get

    OP([HSN]k)=OP([S0(HN)]k)=(1+r)Ntr(ρN[S0(HN)]k)=(1+r)Ntr(Sk0ρN[(Hk)N])=Sk0(1+r)Ntr([ρHk]N)=Sk0(1+r)N(ρ11euk+ρ22euk+ρ33euk+yeu(k1)k+δe(k2)uk(k1)/2)N=Sk0(1+r)N(euk+yeu(k1)k+δe(k2)uk(k1)/2)N=Sk0(1+r)N[1+ak/N+(λa)k/NΔk(k1)/2N+O(1/N2)]NSk0eλe(λ+Δ/2)kk2Δ/2 (38)

    after using the assumptions made on r,u,δ. This provides the limit result (36).

    Remarks. (ⅰ) This representation is useful in computing the European call option OP([HSNK)+]) via the Fourier techniques combined with the Monte Carlo approximation. For the distribution of this process, we view R2 as the complex plane and the planar Brownian motion ~Bt=(B1(t),B2(t)) is then interpreted as a complex-valued Brownian motion.

    (ⅱ) Although observables are traditionally considered to be self-adjoint, the non self-adjoint data is modelled by considering an observable such as H=AB, i.e. the product of two self-adjoint matrices A,B where A and B may represent the non-trade and trade time changes, respectively.

    (ⅲ) Similarly to the convergence to the Black-Scholes model, the limit does not depend on the shift parameter a. However, now the limit depends on the mysterious characteristic Δ.

    In this section, we will apply some methods of quantum mechanics to finance. Discrete and continuous-time markets must be treated separately because of different stochastic behaviors. To simplify the presentation, we assume that the interest rate is 0 and that the risky processes for share prices are martingales.

    The use of Dirac-Feynman quantum mechanics techniques for insurance risk modelling was initiated in Tamturk and Utev (2018) and then developed in Lefèvre et al. (2018); Tamturk and Utev (2019). We adapt this approach to the problem of option pricing in finance, in particular for the pricing of path-dependent options.

    In the Dirac formalism, bra-ket notation is a standard way of describing quantum states. Consider a class of n×n matrices treated as C algebra. A column vector x corresponds to a ket-vector |x>. An associated bra-vector <x| is a row vector defined as its Hermitian conjugate. The usual inner product is denoted by <x|y>, while the outer product |x><y| is the operator/matrix defined by

    |x><y||z>=<y|z>|x>(abc=bca rule). (39)

    In the Feynman path integral methods, the transition probability P(xjxj+1) is computed as the propagator <xj|Aj+1|xj+1> when Aj is a Markovian operator. Thus, the typical path is written as |x0>→|x1>→|xn>, and its probability is given by

    P(|x0>→|x1>→|x2|xn>)=<x0|A1|x1><x1|A2|x2><xn1|An|xn>. (40)

    The main ingredient is then the path calculation formula that calculates the probability P(x0xn) via the sum of the probabilities on all the appropriate paths, i.e.

    P(x0xn)=<x0|A1A2An|xn>=x1,,xn1<x0|A1|x1><x1|A2|x2><xn1|An|xn>. (41)

    It remains to define a suitable propagator for discrete time. For simplicity, we take the operators Aj all equal to A and the time intervals Δti all equal to Δt. In a similar way to e.g. Baaquie (2004); Tamturk and Utev (2018), we assume that the operator A is defined via an Hamiltonian operator H such as A=eΔtH, where H is a Markovian generator called Markovian Hamiltonian. Thus, P(xixi+1)=<xi|eΔtH|xi+1> which can be computed applying the Fourier transform to the momentum space (e.g., Griffiths and Schroeter (2018); Tamturk and Utev (2018)). Specifically, let |p> be a basis in that space, and write <x|p>=eipx and <p|x>=eipx. Then, we get

    <xi|eΔtH|xi+1>=2π0dα2π<xi|eΔtH|α><α|xi+1>=2π0dα2π<xi|α><α|xi+1>eΔtKα=12π2π0(eixiαeixi+1α)eΔtKαdα, (42)

    where {|α>,Kα} is the set of eigenvectors and eigenvalues of the Hamiltonian operator H (i.e., H|α>=Kα|α>). Therefore, we deduce from (41), (42) that

    P(x0xn)=x1,,xn1<x0|eΔtH|x1><x1|eΔtH|x2><xn1|eΔtH|xn>=x1,,xn11(2π)n2π02π02π0[n1i=0(eixiαieixi+1αi)eΔtKαi]dα0dα1dαn1. (43)

    For a more detailed overview of the theory, we refer the reader to the books by Feynman and Hibbs (2010); Griffiths and Schroeter (2018); Parthasarathy (2012) and Plenio (2002).

    Option pricing formula. Consider a claim of the form C=f(S0,S1,...,SN), using the notation of Section 2. From (43), we obtain for the corresponding option price

    OP(C)=dNS0,S1,...,SNf(S0,S1,...,SN)<S0|A1|S1><SN1|AN|SN>. (44)

    The discrete-time approach followed to derive (43) can then be easily applied to the current formula (44).

    A modified Cox-Ross-Rubinstein model. Consider a discrete-time market in which, during each i-th time interval Δt, the share price Si can

    (.) have 1 jump giving Si+1=Sieu,Sied with probabilities p,q=1p, or

    (.) have 2 jumps giving Si+1=Sieu+d,Sie2u,Sie2d with probabilities 2pq,p2,q2, or

    (.) remain the same giving Si+1=Si,

    where d,u are integers with d<0<u. Furthermore, the possible jumps arrive according to a Poisson process of parameter λ so that

    (.) δ1P[N(Δt)=1]=eλΔt(λΔt),

    (.) δ2P[N(Δt)=2]=eλΔt(λΔt)2/2,

    (.) δ0P[N(Δt)=0]1δ1δ2.

    Define xi=ln(Si). The transition probabilities P(SiSi+1) are equivalent to P(xixi+1). Set Δt=1, say. We can now apply (42) with A=eH where the set {|α>,Kα} is found by solving the Schrödinger equation eH|α>=eKα|α> in which

    eH|α>(x)=Ex(eiαxi+1)=eiαx(δ0+eiαupδ1+eiαdqδ1+eiα2up2δ2+eiαx2dq2δ2+eiα(u+d)2pqδ2). (45)

    Notice that the Hamiltonian H is not Markovian, but (42) is still applicable to the discrete times kΔt. Moreover, by construction, the martingale probabilities p=qu and q=qd for the Cox-Ross-Rubinstein model (without interest) yield a martingale in the present situation too since we have

    E(Si+1|Si)=Si[δ0+δ1(queu+qded)+δ2(e2uq2u+e2dq2d+eq+d2quqd)]=Si(δ0+δ1+δ2)=Si. (46)

    We have illustrated numerically the option pricing results obtained for the model. The tables and figures are however too large to be included here.

    This short part is mostly a review of the application of quantum mechanics approach to the continuous time markets and closely follows Baaquie (2004).

    The continuous time formalism is based on the Fourier transform of tempered distributions on the basis |p> in the momentum space. Consider a risky asset price St that evolves in function of an Hamiltonian operator H. First, the method is applied to compute the pricing kernel

    p(x,τ;x)=<x|eτH|x>=dp2π<x|eτH|p><p|x>. (47)

    Then, the option price at time t for the claim QQ(ST), T>t, given St=x is defined by

    OP(Q(ST)|St=x)=<x|e(Tt)H|Q>. (48)

    Black-Scholes model. In this classical approach, the stock price St follows a geometric Brownian motion, i.e. St=S0e˜Bt where ˜Bt=μt+σBt (μ is the drift, σ the volatility and Bt a standard Brownian motion).

    We apply the quantum mechanics approach. Motivated by Baaquie (2004), we work with ˜Bt rather than with St. The corresponding Hamiltonian for the Brownian motion is Hf=(1/2)fμf (computed, for example, via the Itô formula). The Brownian motion kernel is then defined from (47) by the normal density function

    p(u,τ;u)=fN(τμ,τσ2)(uu)=12πτσ2exp[(1/2τσ2)(uuτμ)2]. (49)

    To obtain the option price (48), we set x=S0eu and Q(x)=Q(S0eu)g(u). Then, the Feynman-Kac formula yields

    OP(Q(ST)|St=x)=<u|e(Tt)H|g>=<u|e(Tt)H|u>g(u)du=p(u,Tt;u)Q(S0eu)du. (50)

    Via path integrals. To find the pricing kernel p(x,τ;x)=<x|eτH|x> for τ=Tt, an alternative method consist in using path integral methods. For this, we discretize the time in N intervals of length Δ and consider the xi=x(ti) where ti=iΔ. We then proceed as in the situation in discrete-time. The pricing kernel for (x,x)=(x0,xN) becomes

    p(x,NΔ;x)=dx1dx2dxN1Ni=1<xi1|eΔH|xi>. (51)

    Applications to non-life insurance. Consider Feynman's modification ˜H of the Brownian motion Hamiltonian H by adding the potential V, i.e. for the Hamiltonian ˜H=H+V. For example, choose a ruin level B and take V(x)=+ for x<B. Then, the path calculation formula (43) allows us to compute ruin probabilities when B=0 (Tamturk and Utev (2018, 2019)) and exotic options with barriers when B>0.

    Finally, let us mention that numerically, the binomial model formula (20) and the path integral approach (51) were found to give results close to the Black-Scholes formula, even for relatively small values of N (of order 40).

    We are going to analyse a generated set of financial data by choosing two different quantum models described in Section 2.1. First, we consider the quantum actuarial-type model (see Section 2.1.3 and in 4.1) with the N-step observable defined in (6). Then, we consider the quantum trinomial model (see Section 2.1.2) with the N-step observable defined in (8). Note that matching 1-step for the actuarial case to two steps for the trinomial model is natural, since in the actuarial case, we choose at most two jumps.

    We apply supervised machine learning methods such as developed e.g. in the books by Bishop (2006); Hastie et al. (2009) and Wittek (2014).

    Overall approach. The dataset is supposed to come from a non-ordered class of randomly perturbed observables. Each data is the observation of an eigenvalue λ of the observable perturbed by i.i.d. error terms. The observable is the 1-step operator for the actuarial-type model and the 2-step oparator for the trinomial model. Matching 1-step in the actuarial case to 2-step in the trinomial case is natural since there are at most two jumps in the actuarial case considered.

    First, the data are classified in classes Gλ with respect to the eigenvalues λ. Then, the probabilities pλ are estimated by maximum likelihood using Maxwell-Boltzmann or Bose-Einstein statistics. Finally, the λ are estimated via the weighted L1-norm risk error function.

    Let us explain in more detail for the actuarial-type model, for example.

    Step 1. An initial (u=u0,u=d0) is chosen randomly.

    Step 2. For a given (u,d), the data is classified and labeled against the eigenvalues of the observable using a nearest neighbor algorithm. This leads to the classes Gλ.

    Step 3. For the same (u,d), the estimates ˆp and ˆq are obtained by maximizing the likelihood function L(p,q).

    Step 4. The (u,d) is updated by minimizing a weighted L1-norm risk error function F(u,d)F(λ) defined by

    F(λ)=βλ=λpλβiGλ|βiλ|. (52)

    Step 5. The loop of steps 2 to 4 is repeated until the relative error becomes smaller than a selected difference M, i.e. when

    |F(ui+1,di+1)F(ui,di)|<M. (53)

    kfold cross-validation. To reduce the risk of error, we use a kfold cross-validation strategy. The dataset is randomly divided into k subsets of equal size. One of the subsets is chosen as the training set and the others as test sets. The process is repeated k times, each subset constituting a training element. At each iteration, steps 1 to 5 above are applied to the training data and the results obtained are then checked in the test data. Finally, the estimates used are an average of those obtained on the k iterations.

    Numerical example. As a simple illustration, we will consider the following dataset

    V={86,63,35,52,41,8,24,12,19,24,42,5,91,95,50,49,34,91,37,11}. (54)

    From the assumptions of the model, the 1-step observable H has the eigenvalues

    {λ}={1,ed,eu,ed+u,e2d,e2u}, (55)

    with probabilities respectively given by

    {pλ}={δ0,qδ1,pδ1,pd+uδ2,p2dδ2,p2uδ2}. (56)

    For the probabilities pd+u,q2d,p2u, we consider two possible statistics often used in the analysis of quantum observables.

    Maxwell-Boltzmann independence. This case yields the binomial model since

    pd+u=2pq,p2d=q2,p2u=p2. (57)

    Bose-Einstein dependence. In this case, corresponds to probabilities

    pd+u=Cpq,p2d=Cq2,p2u=Cp2,where C(pq+q2+p2)=1. (58)

    As pointed out in Lefèvre et al. (2018), both statistics admit a formal construction, via the proper choice of the density operator ρ for the number of occurrences and the density projection operators ρλ, such that

    tr(ρρλ)=pλ. (59)

    Likelihood functions. Denote by #x the number of x observed in the data set. For the Maxwell-Boltzmann statistics, the likelihood is defined by the probabilities p, q, δ0, δ1 and δ2 such that

    L(p,q)=L(p,q,δ0,δ1,δ2)=(δ0)#0(qδ1)#d(pδ1)#u(2pqδ2)#u+d(q2δ2)#2d(p2δ2)#2u. (60)

    For the Bose-Einstein statistics, the likelihood function is modified as

    L(p,q)=L(p,q,δ0,δ1,δ2)=(δ0)#0(qδ1)#d(pδ1)#u(Cpqδ2)#u+d(Cq2δ2)#2d(Cp2δ2)#2u. (61)

    Risk functions. From (55), the scaled share price spectrum of HS1 is given by S0{λ}={S0,S0eu,S0ed,S0eu+d,S0e2d,S0e2u}. Thus, using the risk function (52), we get for the Maxwell-Boltzmann case

    F1(u,d)=δ0βiG0|βiS0|+qδ1βiGd|βiS0ed|pδ1βiGu|βiS0eu|+2pqδ2βiGu+d|βiS0eu+d|+q2δ2βiG2d|βiS0e2d|+p2δ2βiG2u|βiS0e2u|, (62)

    and for the Bose-Einstein case,

    F2(u,d)=δ0βiG0|βiS0|+qδ1βiGd|βiS0ed|+pδ1βiGu|βiS0eu|+Cpqδ2βiGu+d|βiS0eu+d|+Cq2δ2βiG2d|βiS0e2d|+Cp2δ2βiG2u|βiS0e2u|. (63)

    Numerical illustration. The data (54) is treated as the eigenvalues of the 1-step observable H defined in (6) for N=1. Let us assume that the Poisson process rate is 1 and the length of time is Δt=1.

    We estimate the values d,u and the probabilities q,p by applying the algorithm of Section 5.1 and using the Maxwell-Boltzmann statistics. Choose, for instance, S0=70, (u0,d0)=(0.8,0.1) and M=0.000001. The results obtained from (60), (62) are presented in Table 1.

    Table 1.  Estimation for N=1 in the Maxwell-Boltzmann case.
    Given (u, d)Maximum likelihoodOptimum (p, q)Optimum (u, d)Risk error F(u,d)|F(ui,di)F(ui+1,di+1)|
    (0.8, -0, 1)4.6063e-14(0.01, 0.99)(0.3, -0.4)43.310743.3107
    (0.3, -0.4)2.2681e-18(0.16, 0.84)(0.3, -0.5)22.873420.4373
    (0.3, -0.5)5.1730e-19(0.23, 0.77)(0.3, -0.6)15.86547.0080
    (0.3, -0.6)1.7631e-18(0.24, 0.76)(0.3, -0.7)14.47411.3913
    (0.3, -0.7)1.2479e-19(0.29, 0.71)(0.3, -0.7)11.19763.2765
    (0.3, -0.7)1.2479e-19(0.29, 0.71)(0.3, -0.7)11.19760

     | Show Table
    DownLoad: CSV

    We then also apply a 4-foldcross-validation procedure. Let Vi, 1i4, before randomly chosen subsets of V such that 4i=1Vi=V. For each iteration, {VVi} and Vi are treated as the training set and the test set, respectively. First, the maximum likelihood estimation and the risk error computation are executed for the training data. Then, the obtained estimates are implemented in the test data. This gives the results of Table 2.

    Table 2.  Using a 4-fold cross-validation strategy.
    Training dataTest data
    Training & Test dataOptimum (u,d)Risk error F(u,d)Optimum (u,d)Risk error F(u,d)Total risk error function
    {VV1},V1(0.2, -0.4)9.5462(0.2, -0.5)0.08429.6304
    {VV2},V2(0.2, -0.3)11.1223(0.2, -0.1)0.192811.3151
    {VV3},V3(0.2, -0.3)8.0269(0.3, -0.4)0.20428.2311
    {VV4},V4(0.2, -0.6)9.2466(0.4, -0.7)0.06299.3095

     | Show Table
    DownLoad: CSV

    We observe that the risk errors are small for the test data but are larger for the training data. Thus, the total risk errors are significant. Similar numerical calculations have also been performed under the Bose-Einstein assumption.

    The spectrum of HN2 is given by

    {λ}={e(Nij)u+id,0jN,0iNj}. (64)

    Thus, for the case N=2, the set of observables is exactly the same as (55). However, the associated probabilities differ from (56) and are equal to

    {pλ}={p22,2p2p3,2p1p2,2p1p3,p23,p21}. (65)

    Likelihood and risk functions. Using the Maxwell-Boltzmann statistics, the likelihood is defined by the probabilities p1, p2, p3 as

    L(p1,p2,p3)=(p22)#0(2p2p3)#d(2p1p2)#u(2p1p3)#d+u(p23)#2d(p21)#2u. (66)

    Note that for the Cox-Ross-Rubinstein model, p2=0 so that the likelihood is simplified to L(p1,p3)=(2p1p3)#d+u(p23)#2d(p21)#2u.

    The corresponding risk function is then defined by

    F3(u,d)=p22βiG0|βiS0|+2p2p3βiGd|βiS0ed|+2p1p2βiGu|βiS0eu|+2p1p3βiGd+u|βiS0eu+d|+p23βiG2d|βiS0e2d|+p21βiG2u|βiS0e2u|. (67)

    Numerical illustration. We process the data (54) again with S0=70, (u0,d0)=(0.8,0.1) and M=0.000001. The results obtained using the algorithm of Section 5.1 with the functions (66), (67) are given in Table 3.

    Table 3.  Estimation for N=2 in the Maxwell-Boltzmann case.
    Given (u, d)Maximum likelihoodOptimum (p1,p2,p3)Optimum (u, d)Risk error F(u,d)|F(ui,di)F(ui+1,di+1)|
    (0.8, -0.1)7.3122e-10(0.01, 0.22, 0.77)(0.5, -0.4)0.03210.0321
    (0.5, -0.4)2.2847e-11(0.1, 0.18, 0.72)(0.8, -0.5)0.32960.2975
    (0.8, -0.5)7.4285e-12(0.1, 0.23, 0.67)(0.8, -0.5)0.33020.0006
    (0.8, -0.5)7.4285e-12(0.1, 0.23, 0.67)(0.8, -0.5)0.33020

     | Show Table
    DownLoad: CSV

    Then, we apply a 4-fold cross-validation procedure as previously done. The obtained results are shown in Table 4. Note that this method allows to reduce the risk errors.

    Table 4.  Using a 4-fold cross-validation strategy.
    Training dataTest data
    Training & Test dataOptimum (u,d)Risk error F(u,d)Optimum (u,d)Risk error F(u,d)Total risk error function
    {VV1},V1(0.3, -0.1)0.1093(0.3, -0.2)0.08600.1953
    {VV2},V2(0.3, -0.1)0.1843(0.7, -0.6)0.00570.1900
    {VV3},V3(0.8, -0.5)0.1591(0.8, -0.6)0.01380.1729
    {VV4},V4(0.7, -0.4)0.2575(0.8, -0.6)0.07210.3296

     | Show Table
    DownLoad: CSV

    The observable operators of the quantum actuarial-type and trinomial models give us different results as expected. Which model to choose? One possible approach might be to consider a mixture of quantum models via the mixture of Hamiltonians (see, e.g., Wittek (2014)).

    Several quantum type financial models are constructed that benefit from the physical interpretation of the unpredictable stock market behaviour and associated dependences. The models provide a general physical type framework for pricing of derivatives and a possibility to construct quantum trading strategies. Moreover, it is revealed that certain quantum type models are applied both in actuarial and financial sciences.

    C. Lefèvre received support from the Chair DIALog sponsored by CNP Assurances. S. Utev benefited from a Scientific Mission granted by the Belgian FNRS. We thank the referees for the careful reading of the paper which greatly helped to improve the presentation.

    All authors declare no conflicts of interest in this paper.



    [1] Baaquie BE (2004) Quantum Finance, Cambridge University Press, Cambridge.
    [2] Baaquie BE (2009) Interest Rates and Coupon Bonds in Qantum Finance, Cambridge University Press, Cambridge.
    [3] Baaquie BE (2014) Path Integrals and Hamiltonians: Principles and Methods, Cambridge University Press, Cambridge.
    [4] Benninga S, Mayshar J (2000) Heterogeneity and option pricing. Rev Deriv Res 4: 7-27. doi: 10.1023/A:1009639211414
    [5] Bishop CM (2006) Pattern Recognition and Machine Learning, Springer, Berlin.
    [6] Black F, Scholes M (1973) The pricing of options and corporate liabilities. J Polit Econ 8: 637-654.
    [7] Bouchaud JP, Potters M (2003) Theory of Financial Risk and Derivative Pricing: From Statistical Physics to Risk Management, 2nd edition, Cambridge University Press, Cambridge.
    [8] Boyle PP (1986) Option valuation using a three-jump process. Int Options J 3: 7-12.
    [9] Chen Z (2001) A non-commutative version of the fundamental theorem of asset pricing. Available from: http://arxiv.org/abs/quant-ph/0112159.
    [10] Chen Z (2004) Quantum theory for the binomial model in finance theory. JSSC. Available from: arXiv preprint quant-ph/0112156.
    [11] Claessens S, Kose MA, Laeven L, et al. (2014)Financial Crises: Causes, Consequences, and Policy Responses, International Monetary Fund, Washington D.C.
    [12] Contreras M, Pellicer R, Villena M, et al. (2010) A quantum model of option pricing: When Black-Scholes meets Schrödinger and its semi-classical limit. Phys A 389: 5447-5459. doi: 10.1016/j.physa.2010.08.018
    [13] Cox JC, Ross SA, Rubinstein M (1979) Option pricing: A simplified approach. J Financ Econ 7: 229-263. doi: 10.1016/0304-405X(79)90015-1
    [14] Feynman RP, Hibbs AR (2010) Quantum Mechanics and Path Integrals, Dover Editions, New York.
    [15] Griffiths DJ, Schroeter DF (2018) Introduction to Quantum Mechanics, 3rd edition, Cambridge University Press, Cambridge.
    [16] Hastie T, Tibshirani R, Friedman J (2009) The Elements of Statistical Learning: Data Mining, Inference, and Prediction, 2nd edition, Springer, New York.
    [17] Haven EE (2002) A discussion on embedding the Black-Scholes option pricing model in a quantum physics setting. Phys A 304: 507-524. doi: 10.1016/S0378-4371(01)00568-4
    [18] Herscovich E (2016) Noncommutative valuation of options. Rep Math Phys 78: 371-386. doi: 10.1016/S0034-4877(17)30015-0
    [19] Karadeniz RS, Utev S (2015) Modelling share prices via the random walk on the lamplighter group. Math Methods Econ Financ 9: 65-84.
    [20] Karadeniz RS, Utev S (2018) Embedding problem for financial data. IIB Int Refereed Acad Soc Sci J 29: 1-18.
    [21] Lefèvre C, Loisel S, Tamturk M, et al. (2018) A quantum-type approach to non-life insurance risk modelling. Risks 6: 1-17. doi: 10.3390/risks6010001
    [22] Leisen DPJ, Reimer M (1996) Binomial models for option valuation-examining and improving convergence. Appl Math Financ 3: 319-346. doi: 10.1080/13504869600000015
    [23] Ma GX, Utev S (2012) Modelling the share prices as a hidden random walk on the lamplighter group, In Mathematical and Statistical Methods for Actuarial Sciences and Finance, Springer, 263-270.
    [24] Mantegna RN, Stanley HE (2000) An Introduction to Econophysics: Correlations and Complexity in Finance, Cambridge University Press, Cambridge.
    [25] Mudakkar SR, Utev S (2013) On stochastic dominance of nilpotent operators, Infinite Dimens Anal Quantum Probab Relat Top, 16: 1350009.
    [26] Parthasarathy KR (2012) An Introduction to Quantum Stochastic Calculus, Birkhäuser, Basel.
    [27] Plenio M (2002) Quantum Mechanics, Ebook, Imperial College, London.
    [28] Tamturk M, Utev S (2018) Ruin probability via quantum mechanics approach. Insur Math Econ 79: 69-74. doi: 10.1016/j.insmatheco.2017.12.009
    [29] Tamturk M, Utev S (2019) Optimal reinsurance via Dirac-Feynman approach. Methodol Comput Appl 21: 647-659. doi: 10.1007/s11009-018-9674-8
    [30] Tian Y (1993) A modified lattice approach to option pricing. J Futures Markets 13: 563-577. doi: 10.1002/fut.3990130509
    [31] Wittek P (2014) Quantum Machine Learning: What Quantum Computing Means to Data Mining, Academic Press, New York.
  • This article has been cited by:

    1. Dominic Cortis, Muhsin Tamturk, Accumulators and Bookmaker’s Capital with Perturbed Stochastic Processes, 2022, 2, 2673-9909, 170, 10.3390/appliedmath2020010
    2. Andrés Gómez, Álvaro Leitao, Alberto Manzano, Daniele Musso, María R. Nogueiras, Gustavo Ordóñez, Carlos Vázquez, A Survey on Quantum Computational Finance for Derivatives Pricing and VaR, 2022, 29, 1134-3060, 4137, 10.1007/s11831-022-09732-9
    3. Muhsin Tamturk, Quantum Computing in Insurance Capital Modelling, 2023, 11, 2227-7390, 658, 10.3390/math11030658
    4. Claude Lefèvre, Muhsin Tamturk, More for Less Insurance Model: An Alternative to (re)Insurance, 2022, 16, 1559-8608, 10.1007/s42519-022-00286-4
    5. Claude Lefèvre, Muhsin Tamturk, Sergey Utev, Marco Carenzo, Cyber Risk in Insurance: A Quantum Modeling, 2024, 12, 2227-9091, 83, 10.3390/risks12050083
    6. Haoran Zheng, Jing Bai, Quantum Leap: A Price Leap Mechanism in Financial Markets, 2024, 12, 2227-7390, 315, 10.3390/math12020315
    7. Youde Dong, Haoran Zheng, Jiehua Zhu, A narrative review on quantum finance theory, 2024, 22, 0219-7499, 10.1142/S0219749924500163
  • Reader Comments
  • © 2019 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(5946) PDF downloads(1120) Cited by(7)

Figures and Tables

Tables(4)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog