Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

Second-order advanced dynamic equations on time scales: Oscillation analysis via monotonicity properties

  • Received: 28 November 2024 Revised: 26 January 2025 Accepted: 17 February 2025 Published: 28 February 2025
  • MSC : 26E70, 34C10, 34K11, 34K42, 34N05

  • This paper derives new oscillation criteria for a class of second-order non-canonical advanced dynamic equations of the form

    (ζ()ϰΔ())Δ+q()ϰ(())=0.

    The derived results are based on establishing dynamic inequalities, which lead to novel monotonicity properties of the solutions. These properties are then used to derive new oscillatory conditions. This approach has been successfully applied to difference and differential equations due to the sharpness of its criteria. However, no analogous studies have adopted a similar methodology for dynamic equations on time scales. Furthermore, this study includes examples to illustrate the importance and sharpness of the main results.

    Citation: Samy E. Affan, Elmetwally M. Elabbasy, Bassant M. El-Matary, Taher S. Hassan, Ahmed M. Hassan. Second-order advanced dynamic equations on time scales: Oscillation analysis via monotonicity properties[J]. AIMS Mathematics, 2025, 10(2): 4473-4491. doi: 10.3934/math.2025206

    Related Papers:

    [1] Min-Ku Lee, Jeong-Hoon Kim . Closed-form approximate solutions for stop-loss and Russian options with multiscale stochastic volatility. AIMS Mathematics, 2023, 8(10): 25164-25194. doi: 10.3934/math.20231284
    [2] Yao Fu, Sisi Zhou, Xin Li, Feng Rao . Multi-assets Asian rainbow options pricing with stochastic interest rates obeying the Vasicek model. AIMS Mathematics, 2023, 8(5): 10685-10710. doi: 10.3934/math.2023542
    [3] Zhidong Guo, Xianhong Wang, Yunliang Zhang . Option pricing of geometric Asian options in a subdiffusive Brownian motion regime. AIMS Mathematics, 2020, 5(5): 5332-5343. doi: 10.3934/math.2020342
    [4] Shoude Huang, Xin-Jiang He . An analytical approximation formula for European option prices under a liquidity-adjusted non-affine stochastic volatility model. AIMS Mathematics, 2022, 7(6): 10364-10377. doi: 10.3934/math.2022577
    [5] Weiwei Shen, Yan Zhang . Strong convergence of the Euler-Maruyama method for the stochastic volatility jump-diffusion model and financial applications. AIMS Mathematics, 2025, 10(5): 12032-12054. doi: 10.3934/math.2025545
    [6] Din Prathumwan, Thipsuda Khonwai, Narisara Phoochalong, Inthira Chaiya, Kamonchat Trachoo . An improved approximate method for solving two-dimensional time-fractional-order Black-Scholes model: a finite difference approach. AIMS Mathematics, 2024, 9(7): 17205-17233. doi: 10.3934/math.2024836
    [7] Shou-de Huang, Xin-Jiang He . Analytical approximation of European option prices under a new two-factor non-affine stochastic volatility model. AIMS Mathematics, 2023, 8(2): 4875-4891. doi: 10.3934/math.2023243
    [8] Min-Ku Lee, Jeong-Hoon Kim . Pricing vanilla, barrier, and lookback options under two-scale stochastic volatility driven by two approximate fractional Brownian motions. AIMS Mathematics, 2024, 9(9): 25545-25576. doi: 10.3934/math.20241248
    [9] Xinyi Wang, Chunyu Wang . Pricing geometric average Asian options in the mixed sub-fractional Brownian motion environment with Vasicek interest rate model. AIMS Mathematics, 2024, 9(10): 26579-26601. doi: 10.3934/math.20241293
    [10] Hijaz Ahmad, Muhammad Nawaz Khan, Imtiaz Ahmad, Mohamed Omri, Maged F. Alotaibi . A meshless method for numerical solutions of linear and nonlinear time-fractional Black-Scholes models. AIMS Mathematics, 2023, 8(8): 19677-19698. doi: 10.3934/math.20231003
  • This paper derives new oscillation criteria for a class of second-order non-canonical advanced dynamic equations of the form

    (ζ()ϰΔ())Δ+q()ϰ(())=0.

    The derived results are based on establishing dynamic inequalities, which lead to novel monotonicity properties of the solutions. These properties are then used to derive new oscillatory conditions. This approach has been successfully applied to difference and differential equations due to the sharpness of its criteria. However, no analogous studies have adopted a similar methodology for dynamic equations on time scales. Furthermore, this study includes examples to illustrate the importance and sharpness of the main results.



    In this paper, we extend the horizon of exotic options by incorporating an activating average condition into the payoff of spread options. As one of the most popular exotic options, spread options are a type of so-called rainbow option whose payoff relies on the price difference (or so-called the spread) between two underlying assets. Spread options are widely traded nowsday both on organized exchanges and over the counter in equity, interest rate, currency, foreign exchange, commodity markets, and energy markets nowadays. For instance, in the energy markets, crack spread options, which either exchange crude oil and unleaded gasoline or exchange crude oil and heating oil, are traded on the New York Mercantile Exchange (NYMEX). They are extensively used to speculate, hedge correlation risks, and even evaluate real assets (see Dempster et al. [23] and Luciano [47]). For a detailed review of different spread option types and their applications, we refer to Carmona and Durrleman [12] and Caldana and Fusai [11].

    Despite their popularity, valuation on spread options written on two underlying assets is an especially challenging problem in quantitative finance. One of the difficulties in pricing spread options is that the exercise boundary is non-linear when the spread is not zero. Obviously, when the spread is zero, this spread option reduces to an exchange option, which allows the holders to exchange one asset for another. Margrabe [48] first deduced the European exchange option pricing formula under the bivariate geometric Brownian motion (GBM) paradigm. Bjerskund and Stensland [6] considered the pricing of American exchange options in the GBM setting. Further analysis and extension to power exchange options is given by Blenman and Clark [5]. The underlying asset pricing model mentioned above assumes that the return of the asset is normally distributed and that its variance is a constant. Recently, there have been a lot of researche on pricing the exchange options under the modified Black-Scholes model (see Black and Scholes [3]) by incorporating with various other factors such as a stochastic interest rate (see Liu and Wang [45]), stochastic volatility (e.g., Antonelli et al. [2], Alos and Rheinlander [1], Kim and Park [35], and Pasricha and Goel [52]), credit risk (e.g., Kim and Koo [33], Wang et al. [61], Pasricha and Goel [51], Xu et al. [66], and Wang et al. [60]), skew-Brownian motion (see Pasricha and He [54]), fractional Brownian motion (e.g., Kim et al. [34] and Kim et al.[36]), jump-diffusion and/or stochastic volatility (e.g., Chen and Wan [17], Cheang and Chaiarella [15], Caldana et al. [10], Wang [58], Li et al. [41], Cufaro-Petroni and Sabino[19], Cheang and Garces [16], Pasricha and Goel [53], and Lian et al. [44]), and so on.

    However, when the spread is not zero, the exercise boundary is non-linear, and it is difficult to obtain an closed-form solutions for these spread options. Instead, we have to resort to analytical approximations or numerical methods. However, practitioners often prefer to use analytical approximations rather than numerical methods because of their computational ease. The pricing issues of the spread options have been investigated in the literature. For instance, Kirk [37] presented an analytical approximation by approximating the sum of the second asset with the fixed strike by a log normal random variable. His method can be thought of as a linear approximation of the exercise boundary. Later, Lo [46] improved Kirk's approximation with an operator splitting method. Pearson [55], Poitras [57], and later Carmona and Durrleman [12], Deng et al. [24], and Bjerksund and Stensland [7] provided lower and upper bounds for the spread option price using suitable approximations of the corresponding discounted expected payoff in log-normal asset models. Caldana and Fusai [11] obtained new lower and upper bounds for the spread option price by using characteristic function and univariate Fourier inversion based on the work of Bjerksund and Stensland [7]. Venkatrmana and Alexander [64] expressed the closed-form price of the spread option as the sum of the prices of two compound exchange options. Kao and Xie [32] proposed a bivariate generalized Edgeworth expansion for pricing spread options. Amongst numerical methods, approaches based on the discrete fast Fourier transform (FFT, see Carr and Madan [13]) have met with large success. For instance, Dempster and Hong [22] introduced a numerical integration method for spread options based on Fourier transforms when the two assets follow Heston's (see Heston [30]) stochastic volatility model. Based on the work of Dempster and Hong [22], there have been many extended results on spread options with models, such as the exponential Levy model (see Hurd and Zhou [31]), GARCH model (see Wang [59]), and Heston stochastic volatility model with jumps (e.g., Olivares and Cane [50] and Hainaut [28]). Furthermore, some researchers have studied other spread option types such as the basket spread option (e.g., Deelstra et al. [20], Pellegrino and Sabono [56], and Lau and Lo [38]) and spread options with credit risk (see Li and Wang [42]).

    As noted above, the payoff at maturity date for the spread option depends on the price at maturity of the two underlying assets alone, which exposes the holder to the risk that the writer may manipulate the two underlying assets' prices such that the payoff of the spread option being benefits according to the writer favorable way near the expiry (see, Deelstra et al. [20]). On the other hand, contracts such as spread options are very common in energy, power, and commodity markets. Especially in energy markets, various forms of average underlying prices are traded, often on the temporal average or multiple underlying assets. Additionally, the average feature can smooth the randomness, or the "noise", inherent in the stock price so that the risk-managers can be evaluated more fundamentally. Therefore, this paper extends the spread options to Asian-spread options using average-rate as the underlying. The payoff of the Asian-spread option depends on the difference between two averages of the underlying asset prices over some predetermined time interval, and has generally the effect of decreasing the variance and offering simpler hedging strategies than regular spread options. Therefore, the price of the spread option combined with the average-rate will be cheaper than that of the regular spread option. That is, this combination would open a wider spectrum of spread payoffs, while making them more accessible to investors or traders at a lower cost. Actually, Asian spread options have recently gained more popularity in the energy market (see, e.g., Carmona and Durrleman [12], Caldana and Fusai [11], Benth and Kruner [4], Deelstra et al. [21], Wang and Zhang [62], and Li et al. [43], and their references therein). The spread part may, for example, be the cost of converting fuel into energy. While the Asian part (the temporal average) avoids the problem common to the European options, namely that speculators can increase gain from the option by manipulating the price of the asset near maturity. The most prominent examples of such contracts is basket spread options, Asian basket spread options, and calendar spread options. These options have been investigated in the options pricing literature. For instance, a multi-asset spread option, such as the basket spread option, resembles the Asian spread option with discrete sampling arithmetic average (see Deelstra et al.[20], Li et al. [40], Pellegrino and Sabino [56], and Lau and Lo [38]) and the moving average exchange options (see, e.g., Han et al. [29]). Choi [18] proposed an efficient and unified method for pricing options such as the basket, spread, and Asian options under multivariate GBM models. A simple, accurate, and efficient method to price and hedge Asian spread options is therefore inevitable.

    The average-rate options, or Asian options, are popular and commonly employed in fields like currency, interest rate, energy, and insurance markets, among others. In general, the average considered in option contracts can be a geometric or arithmetic one, and it can be observed continuously or discretely. Practically, most Asian derivatives on the markets are settled on the arithmetic average price. Usually, the geometric average option can be priced by exact analytical formulas, whereas the arithmetic one does not have closed-form solutions. This is because the probability distribution of the average prices of the underlying asset generally does not have a simple analytical expression. Extensive literature investigates the pricing of the Asian options (see, for example, Levy [39], Castellacci and Siclari [14], Fusai and Kyriakou [26], Willems [63], and Malhotra et al. [49], and their references therein). We refer to Boyle and Boyle [9] for a brief introduction to the development of Asian options.

    In this paper, we propose a theoretical framework for pricing Asian spread options, and derive analytical approximations, which are often preferred to use by spread option traders for their computational ease and the availability of closed-form formulae for hedge ratios. We extend approximation methods of Levy [39], Bjerksund and Stensland [7], and Lau and Lo [38] to the Asian spread case. The main contribution of the present work is twofold. First, we extend spread option to Asian spread options that help those investors who like to mitigate the adverse movements of two underlying assets and hedge both the risk of financial assets over the periods of time. Second, we obtain the derivation of a lower bound, as in Bjerksund and Stensland [7] and Lau and Lo [38], but for spread options with Asian features. Indeed, the only quantity we need to know explicitly is the joint probability density function of the log-returns of the two averages.

    The remainder of this paper is organized as follows: Section 2 describes the market model and the Asian spread options. In Section 3, the closed-form approximate formulas of the spread options with geometric averaging and arithmetic averaging are derived respectively. Numerical examples are presented in Section 4 to show the accuracy and efficiency of the proposed method, while the effects of some parameters on options and their deltas are analyzed. Finally, Section 5 concludes this paper.

    Let (Ω,F,(F)t0,Q) be a complete probability space equipped with a filtration (Ft)t0 satisfying the usual conditions. Moreover, denote by E() the expectation operator with respect to a risk neutral equivalent martingale measure Q. A continuous-time financial market is considered with a finite time horizon [0,T], where T< under the complete probability space (Ft)t[0,T]. Assume that there exists two risky assets and the risk-free asset traded continuously in this financial market over a finite time interval [0,T]. Let the process of the risk-free asset is governed by

    dBt=rBtdt,B0=1,

    where r is the risk-free interest rate. Under the risk neutral probability Q, the two risky assets whose prices are denoted by S1t and S2t are governed by the following stochastic differential equations

    dS1tS1t=rdt+σ1dW1t, (2.1)
    dS2tS2t=rdt+σ2dW2t, (2.2)

    where σis(i=1,2) are the volatilities of both assets. In addition, W1t and W2t are two standard Brownian motions defined on this filtered probability space. We assume that the correlation coefficient between W1t and W2t is given by ρ. Assume that r,σ1,σ2 and ρ are constants, and (Ft)t0 is produced by the σ-algebra of the price pair (S1t,S2t)t0.

    Now, we present the payoff of Asian spread options (ASO). The payoff of these options is based on the difference between two average asset prices of two underlying assets for the period up to T. In the following, we introduce the payoff function of the Asian spread options as follows:

    h(x1,x2,K)=(x1x2K)+, (2.3)

    where xi can be either GiT or AiT depending on the geometric average price or arithmetic average price, respectively. The notation x+=max{x,0} and K0 is the strike price of this option. Additionally, Git(i=1,2) is the continuously monitored geometric average of Siu(i=1,2) over time [0,t], that is,

    Git=exp(1tt0lnSiudu),

    and Ait(i=1,2) is the continuously monitored arithmetic average of Siu over time [0,t], i.e.,

    Ait=1tt0Siudu.

    There is no known closed form for this case defined in (2.3) despite the use of the more tractable geometric average.

    From (2.1) and (2.2), we have for any t[0,T],

    lnSit=lnSi0+(r12σ2i)t+σiWit,i=1,2,

    and

    lnGiT=1TT0lnSitdt=lnSi0+(r12σ2i)T2+σiTT0(Tt)dWit. (2.4)

    Accordingly, for any constant α, one gets that

    E(GαiT)=Sαi0exp{α(r12σ2i)T2+α26σ2iT}, (2.5)

    and from the result of Geman and Yor [27],

    E(AiT)=Si0(erT1)rT, (2.6)
    E(A2iT)=2S2i0[re(2r+σ2i)T(2r+σ2i)erT+(r+σ2i)]rT2(r+σ2i)(2r+σ2i),i=1,2. (2.7)

    In the following, we investigate the pricing problem for the Asian spread options under the payoff functions in the geometric average and arithmetic average cases described previously.

    Without loss of generality, we focus on the option price at the inception (t=0). Let GASO(S1,S2,K) be the price at time t=0 for the ASO whose payoff function is h(G1T,G2T,K), and AASO(S1,S2,K) be the price at time t=0 for the ASO whose payoff function is h(A1T,A2T,K). According to the risk-neutral pricing theory, the option prices are thus given by

    GASO(S1,S2,K)=E{erT(G1TG2TK)+}, (3.1)
    AASO(S1,S2,K)=E{erT(A1TA2TK)+}, (3.2)

    respectively. Obviously, a more difficult problem for pricing the Asian spread options defined in the above work is the unknown joint distribution of two arithmetic average prices (A1T,A2T) and the non-linear exercise boundary with the spread K being not zero. In this paper, we will use approximation approaches to solve this problem. The next result is well-known.

    Proposition 1. Assume that random variables X1 and X2 satisfy X1N(μ1,δ21), X2N(μ2,δ22), and ϱ=Corr(X1,X2). In addition, a,b,c,d,e,μ1,μ2,δ1,δ2 and ϱ are assumed to be constants, where at least one of c and d is non-zero. Then,

    E[eaX1+bX21(cX1+dX2e)]=exp[aμ1+bμ2+12(a2δ21+2ϱabδ1δ2+b2δ22)]N(cμ1+dμ2e+acδ21+ϱ(ad+cb)δ1δ2+bdδ22c2δ21+2ϱcdδ1δ2+d2δ22), (3.3)

    where N() denotes the standard normal cumulative distribution function and 1A is an indicator function for any event A.

    Proof: It is obvious that (aX1+bX2)(aμ1+bμ2)a2δ21+2ϱabδ1δ2+b2δ22N(0,1) and (cX1+dX2)(cμ1+dμ2)c2δ21+2ϱcdδ1δ2+d2δ22N(0,1). Then, it follows from the lemma in Dravid et al. [25] that the proof is completed.

    In order to value the Asian spread options defined in (2.3), we first requires to determine the linear exercise boundary in logarithmic variables from the exercise region {G1TG2TK} (or {A1TA2TK}). In this paper, along with the method used in Bjerksund and Stensland [7] for options written on the spread between two assets, and Lau and Lo [38] for multi-asset basket spread options, we derive an approximated closed-form formula by modifying the origin exercise region slightly. Second, we need to know the joint probability distribution of two arithmetic average prices. Here, we will apply distribution-approximating and moment-matching (see, e.g., Brignone et al. [8]) methods to derive the joint probability distribution and extend the work in Levy [39] from one-dimensional unknown arithmetic average distribution by the corresponding log-normal distribution to two-dimensional cases. Actually, our approximated formula is always a little less than the fair value of spread options so that it can be seen as a lower bound. Now, we are in the position to state the main theoretical results of this paper.

    Proposition 2. Based on the proposed model specification (2.1) and (2.2), the price at time t=0 for the ASO on geometric average is given by

    GASO(S1,S2,K)=S10e(r+16σ21)T2N(d1)S20e(r+16σ22)T2N(d2)KerTN(d3), (3.4)

    where di,i=1,2,3 are given by

    d1=lnS10k+[(r+16σ21)12+16α2σ2213ρασ1σ2]TσT,d2=lnS10k+[16α2σ2213ασ22+(r12σ21)12+13ρσ1σ2]TσT,d3=lnS10k+[(r12σ21)12+16α2σ22]TσT,

    in which the parameters σ,α and k are

    σ=13(σ212ρασ1σ2+α2σ22),α=E(G2T)E(G2T)+K,k=E(G2T)+K.

    Proof: See the Appendix.

    Proposition 3. Based on the proposed model specification (2.1) and (2.2), the price at time t=0 for the ASO on arithmetic average is given by

    AASO(S1,S2,K)=erT[eμ1+12δ21N(ˆd1)eμ2+12δ22N(ˆd2)KN(ˆd3)], (3.5)

    where ˆdi,i=1,2,3 are given by

    ˆd1=μ1ˆk+(δ21ˆρβδ1δ2+12β2δ22)δ212ˆρβδ1δ2+β2δ22,ˆd2=μ1ˆk+(ˆρδ1δ2βδ22+12β2δ22)δ212ˆρβδ1δ2+β2δ22,ˆd3=μ1ˆk+12β2δ22δ212ˆρβδ1δ2+β2δ22,

    in which the parameters above are

    β=E(A2T)E(A2T)+K,ˆk=ln[E(A2T)+K],μi=2lnE(AiT)12lnE(A2iT),δ2i=lnE(A2iT)2lnE(AiT),i=1,2,ˆρ=2[lnE(A1TA2T)(μ1+μ2)](δ21+δ22)2δ1δ2,E(A1TA2T)=2S10S20rT2[e(2r+ρσ1σ2)TerT(r+ρσ1σ2)e(2r+ρσ1σ2)T1(2r+ρσ1σ2)].

    Proof: See the Appendix in detail.

    Remark 1. First, the formulas (3.4) and (3.5) are the lower bounds of the exact Asian spread option prices, and in practice they are so tight that they could be seen as an accurate approximation of the true value. These will be shown through numerical experiments in Section 4. Second, the values of parameter α (or β) and k (or ˆk) given in the above expressions are proposed by Bjerksund and Stensland [7] in the log-normal setting within the underlying assets. They are also effective when we take average-rate prices into consideration (see Lau and Lo [38] for multi-asset basket spread options whose payoff is the linear combination of several underlying assets). The above formula is very easy to implement by the software such as Matlab. In the numerical section, we shall show the accuracy and efficiency of the approximation pricing formulas.

    Remark 2. When K<0 by using the fact that E[X+]=E[(X)+]+E[X], we can obtain that

    E{erT(G1TG2TK)+}=E{erT[(G2TG1T˜K)+]}=E{erT(G2TG1T˜K)+}E{erT(G2TG1T˜K)}=GASO(S2,S1,K)S20exp{12(r+16σ22)T}+S10exp{12(r+16σ21)T}KerT,

    and

    E{erT(A1TA2TK)+}=AASO(S2,S1,K)S20(1erT)rT+S10(1erT)rTKerT,

    where ˜K=K>0.

    A special case of the above formula has been extensively studied. For example, if we set σ2=0 and ρ=0, we have an Asian option. In addition, the formulas (3.4) and (3.5) have two simple forms when K=0. In fact, when K=0, the formulas (11) and (12) reduce to the pricing formulas for the Asian exchange options which are similar to the works by Margrabe [48] and Han et al. [29].

    Corollary 1. If we let K=0 in (11) and (12), we obtain the following formulas for the Asian exchange options:

    GAEO(S1,S2)=S10e(r+16σ21)T2N(lnS10S20+(r+112σ2114σ22)TT3(σ212ρσ1σ2+σ22))S20e(r+16σ22)T2N(lnS10S20(14σ21+112σ2213ρσ1σ2)TT3(σ212ρσ1σ2+σ22)), (3.6)

    and

    AAEO(S1,S2)=erT[eμ1+12δ21N(μ1μ2+12(δ21ˆρδ1δ2)δ212ˆρδ1δ2+δ22)eμ2+12δ22N(μ1μ2+12(ˆρδ1δ2δ22)δ212ˆρδ1δ2+δ22), (3.7)

    respectively.

    Proof. When K=0, then α=1, β=1, k=E(G2T), and ˆk=lnE(A2T) defined in formulas (3.4) and (3.5). In this case, k=S20exp{(r16σ22)T2} and ˆk=μ2+12δ22. Hence, it follows that (3.6) and (3.7) hold.

    The approximation can be applied to the Greeks computation, as well. In particular, we can derive the explicit expressions for the delta, which is defined as the rate of change of the option value with respect to the underlying asset price, of the ASO priced by our two approximations, GASO and AASO. Similar formulas can be computed for the other Greeks, such as Gamma, Vega, Theta, and Rho. The following corollary summarizes our findings.

    Corollary 2. The hedging parameter Δ of the ASO, based on the GASO and AASO pricing formulas, are respectively

    Δ(g)1=GASOS1=e(r+16σ21)T2N(d1)+e(r+16σ21)T2d212σ2πTS2e(r+16σ22)T2d222S1σ2πTKerTd232S1σ2πT,Δ(g)2=GASOS2=αS1e(r+16σ21)T2d212S2σ2πT{1+[1+d1σT](1α)(ασ22ρσ1σ2)T3}e(r+16σ22)T2N(d2) (3.8)
    +αe(r+16σ22)T2d222σ2πT[1+(1α)23σ22T+(1α)(ασ22ρσ1σ2)d2T3σ]+αKe(rT+d232)S2σ2πT[1α(1α)3σ22T+(1α)(ασ22ρσ1σ2)d3T3σ], (3.9)
    Δ(a)1=AASOS1=erT{1S1eμ1+12δ21N(ˆd1)+S12πδ[eμ1+12(δ21ˆd21)(1ˆd1[11S21]δ)+eμ2+12(δ22ˆd22)(12S21+ˆd2[11S21]δ)+Ke12ˆd23(12S21+ˆd3[11S21]δ)]}, (3.10)
    Δ(a)2=AASOS2=erTS2{eμ2+12δ22N(ˆd2)+β2πδ[eμ1+12(δ21ˆd21)(1[β(S221)ˆρ(1β)δ1δ2+β(1β)δ22][1ˆd1δ])+eμ2+12(δ22ˆd22)(1+[1β][(S221)+(1β)δ22]+ˆd2δ[β(S221)+β(1β)δ22ˆρ(1β)δ1δ2])+Ke12ˆd23(1β[(S221)+(1β)δ22]+ˆd3δ[β(S221)+β(1β)δ22ˆρ(1β)δ1δ2])]}, (3.11)

    where δ=δ212ˆρβδ1δ2+β2δ22.

    Proof: Using the results of Propositions 2 and 3, straightforward calculations lead to the formulas (3.8)–(3.11).

    In this section, we present a few numerical examples of our main results in Propositions 2 and 3 above. We first compare the performance of Propositions 2 and 3 with the Monte Carlo (MC) simulation in Table 1 below. The MC simulation values were used for the benchmark, and accuracy was measured by the root of mean-squared errors (RMSE) and maximum absolute error (MAE). In our comparison, we continued to use the same model parameters as the numerical examples in [7]. More specifically, we picked the following model parameters in our numerical tests: r=0.05,S10=100,S20=96,σ1=0.2, and σ2=0.1. The MC simulation was implemented by means of the Euler-Maruyama discretization method. In our numerical experiments, we generated N=1,000,000 sample paths with running on a daily basis. Second, we compared the option prices for the ASOs with those of the commom spread options. Finally, in order to see the impact of underlying parameters on approximated prices and delta values, sensitivity analysis was conducted. All the algorithms were implemented in MatLab (R2018b). The codes for the examples were run in MATLAB R2017a on a PC with the configuration: Intel(R)Core(TM) i7-8550UCPU@1.80GHz 1.99GHz and 8.0GB RAM.

    Table 1.  Accuracy comparison of pricing formulas of ASO for the case T=1(year).
    ρ K GASO GMC 95% C.I. AASO AMC 95% C.I.
    -0.5 0.0 7.8154 7.7813 (7.7550, 7.8176) 8.0166 7.9646 (7.9377, 8.0916)
    0.4 7.5943 7.5850 (7.5589, 7.6111) 7.7933 7.7667 (7.7400, 7.7934)
    0.8 7.3771 7.3602 (7.3346, 7.3859) 7.5740 7.5417 (7.5155, 7.5780)
    1.2 7.1638 7.1871 (7.1617, 7.2125) 7.3585 7.3671 (7.3411, 7.3932)
    1.6 6.9545 6.9524 (6.9272, 6.9775) 7.1470 7.1317 (7.1059, 7.1575)
    2.0 6.7492 6.7358 (6.7111, 6.7605) 6.9393 6.9129 (6.8876, 6.9403)
    2.4 6.5478 6.5173 (6.4930, 6.5486) 6.7356 6.6926 (6.6677, 6.7376)
    2.8 6.3504 6.3386 [6.3145, 6.3627) 6.5358 6.5136 (6.4888, 6.5383)
    3.2 6.1569 6.1572 (6.1335, 6.1809) 6.3400 6.3299 (6.3056, 6.3543)
    3.6 5.9674 5.9611 (5.9377, 5.9846) 6.1480 6.1328 (6.1088, 6.1569)
    4.0 5.7819 5.7969 (5.7737, 5.8200) 5.9600 5.9673 (5.9435, 5.9911)
    0.0 0.0 6.9506 6.9027 (6.8798, 6.9555) 7.1500 7.0812 (7.0578, 7.1546)
    0.4 6.7239 6.7127 (6.6900, 6.7353) 6.9209 6.8900 (6.8668, 6.9232)
    0.8 6.5019 6.4958 (6.4735, 6.5181) 6.6963 6.6719 (6.6490, 6.6988)
    1.2 6.2845 6.2959 (6.2739, 6.3178) 6.4763 6.4711 (6.4485, 6.4936)
    1.6 6.0717 6.0639 (6.0423, 6.0855) 6.2609 6.2372 (6.2149, 6.2694)
    2.0 5.8637 5.8213 (5.8000, 5.8726) 6.0502 5.9939 (5.9720, 6.0558)
    2.4 5.6603 5.6620 (5.6409, 5.6831) 5.8441 5.8320 (5.8103, 5.8537)
    2.8 5.4616 5.5168 (5.4960, 5.5376) 5.6426 5.6879 (5.6665, 5.7093)
    3.2 5.2676 5.3129 (5.2924, 5.3333) 5.4457 5.4820 (5.4609, 5.5031)
    3.6 5.0783 5.0836 (5.0635, 5.1037) 5.2535 5.2491 (5.2284, 5.2699)
    4.0 4.8936 4.9114 (4.8916, 4.9311) 5.0659 5.0757 (5.0553, 5.0961)
    0.5 0.0 5.9065 5.9083 (5.8896, 5.9270) 6.1050 6.0865 (6.0673, 6.1058)
    0.4 5.6695 5.6527 (5.6343, 5.6710) 5.8649 5.8273 (5.8084, 5.8663)
    0.8 5.4384 5.4560 (5.4377, 5.4742) 5.6307 5.6294 (5.6106, 5.6482)
    1.2 5.2133 5.2160 (5.1981, 5.2339) 5.4023 5.3861 (5.3676, 5.4045)
    1.6 4.9942 4.9812 (4.9637, 4.9987) 5.1798 5.1509 (5.1328, 5.1890)
    2.0 4.7811 4.7857 (4.7684, 4.8029) 4.9633 4.9523 (4.9345, 4.9702)
    2.4 4.5740 4.5887 (4.5718, 4.6056) 4.7527 4.7525 (4.7350, 4.7700)
    2.8 4.3729 4.3362 (4.3197, 4.3827) 4.5480 4.4977 (4.4806, 4.5547)
    3.2 4.1779 4.1472 (4.1311, 4.1833) 4.3493 4.3057 (4.2890, 4.3524)
    3.6 3.9888 3.9790 (3.9632, 3.9949) 4.1566 4.1354 (4.1189, 4.1578)
    4.0 3.8057 3.8129 (3.7973, 3.8285) 3.9697 3.9691 (3.9529, 3.9853)
    RMSE 0.0222 0.0298
    MAE 0.0552 0.0688
    Time (se) 0.0192 246.46 0.0273 258.75

     | Show Table
    DownLoad: CSV

    Table 1 provides the computational results for 33 ASOs with different strike prices. In Table 1, GASO and AASO are the values obtained by our pricing formulas (3.4) and (3.5), and 'GMC' and 'AMC' are the values by MC simulations for the ASO on geometric average and arithmetic average, respectively. The columns labeled "95% C.I." is the 95% confidence interval for the MC simulation method. From Table 1, we can see that the approximated formulas in Propositions 2 and 3 gave highly accurate Asian spread option prices as they were very close to the prices obtained by the MC simulation. Concretely, the RMSE was less than 3% and the MAE was less than 7% for all cases. In addition, we observed that the speed of this approximated method was faster than the MC simulation method. Concretely, the average of the computational time of the MC methods was over 8 seconds, while the approximated approach took less than 0.008 second. As a result, the approximated formulas in Propositions 2 and 3 were pretty tight so that they could be used as an excellent approximation.

    Table 2 reports comparisons between the ASOs prices and the plain-vanilla spread option prices derived by Bjerksund and Stensland [7]. As expected, it can be seen from Table 2 that the ASOs prices were all less than those of the plain-vanilla spread option, and the prices of the ASO with arithmetic averaging were higher than those of the ASO with geometric averaging under the same parameter values. In addition, higher strike prices K and higher correlation coefficients ρ led to lower Asian spread call option prices. This finding was similar to that for the plain-vanilla spread option in Hurd and Zhou [31], Bjerksund and Stensland [7], and Lo [46]. On the other hand, we also observed that the prices of the spread options including the ASOs and plain-vanilla spread option increased as the maturity time T increased, meaning that the maturity had a significant effect on the values of Asian spread options. Intuitively, the values of Asian spread options depended on the relative performances between two underlying assets. This may be explained by the fact that the volatilities of the two underlying assets in a long time period were larger, so the option price rose.

    Table 2.  Comparison of ASOs and spread options.
    T=1 T=3
    ρ K GASO AASO Spread option GASO AASO Spread option
    -0.5 0.0 7.8154 8.0166 12.4356 11.0883 11.7499 19.8298
    0.4 7.5943 7.7933 12.2317 10.9077 11.5660 19.6595
    0.8 7.3771 7.5740 12.0301 10.7290 11.3841 19.4903
    1.2 7.1638 7.3585 11.8307 10.5524 11.2041 19.3221
    1.6 6.9545 7.1470 11.6336 10.3778 11.0261 19.1551
    2.0 6.7492 6.9393 11.4387 10.2052 10.8500 18.9891
    2.4 6.5478 6.7356 11.2461 10.0345 10.6759 18.8242
    2.8 6.3504 6.5358 11.0557 9.8659 10.5037 18.6604
    3.2 6.1569 6.3400 10.8676 9.6993 10.3334 18.4977
    3.6 5.9674 6.1480 10.6817 9.5346 10.1651 18.3361
    4.0 5.7819 5.9600 10.4981 9.3720 9.9987 18.1756
    0.0 0.0 6.9506 7.1500 10.8684 9.6365 10.2755 17.1303
    0.4 6.7239 6.9209 10.6620 9.4544 10.0895 16.9603
    0.8 6.5019 6.6963 10.4583 9.2745 9.9058 16.7916
    1.2 6.2845 6.4763 10.2572 9.0971 9.7244 16.6242
    1.6 6.0717 6.2609 10.0589 8.9221 9.5453 16.4581
    2.0 5.8637 6.0502 9.8632 8.7494 9.3686 16.2933
    2.4 5.6603 5.8441 9.6702 8.5791 9.1941 16.1298
    2.8 5.4616 5.6426 9.4798 8.4111 9.0220 15.9675
    3.2 5.2676 5.4457 9.2921 8.2455 8.8521 15.8066
    3.6 5.0783 5.2535 9.1071 8.0822 8.6845 15.6469
    4.0 4.8936 5.0659 8.9247 7.9213 8.5192 15.4885
    0.5 0.0 5.9065 6.1050 8.9497 7.8549 8.4637 13.7923
    0.4 5.6695 5.8649 8.7386 7.6698 8.2737 13.6229
    0.8 5.4384 5.6307 8.5311 7.4877 8.0867 13.4552
    1.2 5.2133 5.4023 8.3269 7.3088 7.9027 13.2891
    1.6 4.9942 5.1798 8.1263 7.1329 7.7217 13.1248
    2.0 4.7811 4.9633 7.9290 6.9601 7.5436 12.9621
    2.4 4.5740 4.7527 7.7352 6.7903 7.3685 12.8011
    2.8 4.3729 4.5480 7.5449 6.6235 7.1964 12.6417
    3.2 4.1779 4.3493 7.3579 6.4598 7.0272 12.4839
    3.6 3.9888 4.1566 7.1743 6.2991 6.8609 12.3278
    4.0 3.8057 3.9697 6.9942 6.1413 6.6975 12.1733

     | Show Table
    DownLoad: CSV

    The impacts of basic parameters on the ASOs prices or their deltas are shown in Figures 13, including the initial values and volatilities of the underlying assets. Figure 1 shows that a larger S2 results in a larger option price with other parameters being fixed, but the impact of the parameter S1 on the ASOs price is opposite with a larger S1 corresponding to a lower option price, where other parameters are fixed. In other words, a larger difference (S1S2) based on the two underlying asset prices results in a lower price for the ASOs.

    Figure 1.  Prices of ASOs against different initial values S1 and S2.
    Figure 2.  Impact of option price for different volatilities σ1 and σ2.
    Figure 3.  Δ values against different initial values S1 or S2.

    Figure 2 depicts the influence of the volatilities σ1 and σ2 of the two underlying assets. The prices of the ASO options increase with the volatilities. This increases the volatility of the spread and hence the spread option value.

    Figure 3 displays the Δ values of the GASO and AASO against the underlying asset prices S1 or S2. From Figure 3(a), one can observe that the Δ values of GASO increase as S1 increases, and become larger as the strike price K decreases. Contrary to the case of the underlying asset price S2, the Δ values of GASO decrease as S2 increases and become larger as the strike price K increases. Also, there is a similar pattern for the Δ values of AASO with respect to S2 and K from Figure 3(b).

    Figure 4 illustrates the changes in the Δ value of GASO, AASO, and the spread option with respect to S1 or S2, respectively. We observed that an increase in the initial underlying asset value led to a rise in the option's delta value. However, the delta value increased at a slower pace for the spread option compared to the Asian spread options. This is a result of the average value of the underlying asset over the entire period, which affected a portion of the option's delta value and made it less sensitive than in the case where it was based on the underlying asset's value at the expiration date.

    Figure 4.  Comparison of Δ values against different initial values S1 or S2.

    This paper presents a new variety of financial instruments-spread option on two average asset prices. In virtue of the moment-matching and distribution-approximating techniques, we generalized the Bjerksund and Stensland [7] approximate of two asset spread option formulas to the case of the Asian spread options, and obtained the analytical valuation formulas for the Asian spread options on geometric averaging and arithmetic averaging. Finally, we presented some numerical results. The main contribution of this paper is to provide practitioners with a pricing formula, which can be used for real-time pricing of Asian spread options. For example, practitioners might apply the above option pricing formulas (3.4) and (3.5) to calibrate the model and estimate the model parameters on a set of market data of European-style Asian spread options by minimizing the difference between market prices and model prices in a least-squared error fit.

    At present, although trading of the options proposed in this paper is very limited in the real market, financial innovations have become increasingly important to risk management. Creating an option is unusual and is more likely to occur when academics are involve. Therefore, the Asian spread options that non-trivially combine the spread option with Asian options obviously provide great value-added potentials in financial markets. In particular, Asian spread options not only have wide applications in risk management but also appropriately enter the payoff function of incentive contracts for management compensation.

    There are a few directions that one can take to extend and improve the results in this paper. First, in the geometric Brownian motions case, our results can be easily extended to incorporate jumps or stochastic volatility (for example, the Merton jump-diffusion model or the Hull-White stochastic volatility model) in the price processes of the assets. Second, the approximation method might be improved for Asian spread options such as using the Edgeworth expansion approach by Kao and Xie [32], or the method considered in Castellaccia and Siclari [14], who approximate an arithmetic average with a geometric one by adjusting the strike for the discrepancy. In addition, the formulas of continuously monitored average Asian spread options can be a good proxy for the corresponding discrete-type option in cases that are a kind of basket option. We leave these interesting topics for future research.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    This work was supported by the NSF of China (grant number: 11461008), Guangxi Natural Science Foundation (grant number: 2018GXNSFAA281016), and the Project for enhancing Young and Middle-aged teacher's research basic ability in colleges of Guangxi (grant number: 2021KY0588).

    The authors thank the editors and the anonymous referees for their valuable suggestions, which have significantly improved this paper. The authors are responsible for the remaining errors.

    The authors declare that there are no conflicts of interest regarding the publication of this paper.

    Proof of Proposition 2: We define the following event,

    A={ω:G1TGα2T>ekE(Gα2T)}, (A.1)

    and then from (2.4) and (2.5)

    A={ω:S10e(r12σ21)T2+σ1TT0(Tt)dW1tSα20e(r12σ22)αT2+ασ2TT0(Tt)dW2t>ekE(Gα2T)}={ω:σ1TX1ασ2TX2>lnkS1016α2σ22T(r12σ21)T2}, (A.2)

    where X1=T0(Tt)dW1tN(0,T33) and X2=T0(Tt)dW2tN(0,T33). Therefore, following the idea of Bjerksund and Stensland [7], and [38], we have (G1TG2TK)+(G1TG2TK)1A, and get a lower bound to the Asian spread options

    GASO(S1,S2,K)=E{erT(G1TG2TK)1A}=erT(I1I2I3),

    where I1=E{G1T1A},I2=E{G21T1A} and I3=KE{1A}. In the following, we adopt Proposition 1 to derive the expressions of I1I3 in turn. First, we derive I1 as follows:

    I1=E{G1T1A}=E[S10e(r12σ21)T2+σ1TX11A]=E[S10e(r12σ21)T2+σ1TX11(σ1TX1ασ2TX2>lnkS1016α2σ22T(r12σ21)T2)]=S10e(r16σ21)T2N(d1). (A.3)

    Similarly, it holds that

    I2=E{G2T1A}=E[S20e(r12σ22)T2+σ2TX21A]=E[S20e(r12σ22)T2+σ2TX21(σ1TX1ασ2TX2>lnkS1016α2σ22T(r12σ21)T2)]=S20e(r16σ22)T2N(d2), (A.4)

    and

    I3=KE{1A}=KQ(σ1TX1ασ2TX2>lnkS1016α2σ22T(r12σ21)T2)=KN(d3). (A.5)

    Now, we have completed the proof.

    Proof of Proposition 3: Due to the fact that the integral of lognormal distribution no longer obeys lognormal distribution, its probability distribution is difficult to determine. In the following, we adopt the wilkinson approximation of Levy [39] for one-dimensional lnAT with a two-parameter lognormal distribution to extend the two-dimensional case (lnA1T, lnA2T) with five-parameter adjoint lognormal distribution approximation. In this case, we assume that

    (lnA1T,lnA2T)N2(μ1,μ2,δ21,δ22,ˆρ), (A.6)

    where N2() denotes the bivariate normal cumulative distribution function. It is obvious that the moment generating function for the two-dimensional random vector (lnA1T,lnA2T) is

    E(eλ1lnA1T+λ2lnA2T)=exp{λ1μ1+λ2μ2+12(λ21δ21+2ˆρλ1λ2δ1δ2+λ22δ22)},

    and it holds then that

    E(elnA1T)=E(A1T)=exp{μ1+12δ21},E(e2lnA1T)=E(A21T)=exp{2μ1+2δ21},E(elnA2T)=E(A2T)=exp{μ2+12δ22},E(e2lnA2T)=E(A22T)=exp{2μ2+2δ22},E(elnA1T+lnA2T)=E(A1TA2T)=exp{μ1+μ2+12(δ21+2ˆρδ1δ2+δ22)}.

    Therefore, we have the five parameters as follows:

    μi=2lnE(AiT)12lnE(A2iT),δ2i=lnE(A2iT)2lnE(AiT),i=1,2,ˆρ=2[lnE(A1TA2T)(μ1+μ2)](δ21+δ22)2δ1δ2,

    where E(AiT) and E(A2iT) for i=1,2 are defined by (2.6) and (2.7). Now, it is key to compute the expectation E(A1TA2T) in the third expression above.

    Next, we use the theory of the polynomial diffusion process suggested by Willems [63] to compute such moment E(A1TA2T). Since

    tAit=Si0t0e(r12σ2i)u+σiWiudulaw=Si0t0e(r12σ2i)(tu)+σi(WitWiu)du:=Si0Zit, (A.7)

    for i=1,2 and fixed t>0. Using Ito's formula, we have

    dZit=(1+rZit)dt+σiZitdWit,i=1,2, (A.8)

    and Zit,i=1,2 are polynomial diffusion processes. So, we know that E(A1TA2T)=S10S20T2E(Z1TZ2T). On the other hand, we get

    d(Z1tZ2t)=Z1tdZ2t+Z2tdZ1t+dZ1tdZ2t=(Z1t+Z2t+2rZ1tZ2t+ρσ1σ2Z1tZ2t)dt+Z1tZ2t(σ1dW1t+σ2dW2t).

    Hence, we have (using Fubini's theorem)

    E(Z1tZ2t)=t0[E(Z1u)+E(Z2u)+(2r+ρσ1σ2)E(Z1uZ2u)]du, (A.9)

    where E(Ziu)=ert1r,i=1,2. it shows that y(t)=E(Z1tZ2t) is the solution of the following ordinary differential equation

    {dy(t)dt(2r+ρσ1σ2)y(t)=2ert1r,y(0)=0. (A.10)

    After solving the equation above, we obtain:

    E(Z1tZ2t)=2r[e(2r+ρσ1σ2)tertr+ρσ1σ2e(r+ρσ1σ2)t12r+ρσ1σ2]. (A.11)

    Finally, similar to the proof of Proposition 2, we calculate the price of the AASO as follows:

    B={ω:A1TAβ2T>eˆkE(Aβ2T)}={ω:lnA1TβlnA2T>ˆkβμ212β2δ22}. (A.12)

    Thus, the price of the AASO is given by

    AASO(S1,S2,K)=E{erT(A1TA2TK)1B},=erT(ˆI1ˆI2ˆI3),

    where

    ˆI1=E(A1T1B)=E[elnA1T1(lnA1TβlnA2T>lnD2)]=eμ1+12δ21N(ˆd1),ˆI2=E(A2T1B)=E[elnA2T1(lnA1TβlnA2T>lnD2)]=eμ2+12δ22N(ˆd2),ˆI3=KE(1B)=KQ(lnA1TβlnA2T>lnD2)=KN(ˆd3).

    By combining the expressions above, we obtain the stated formulae.



    [1] S. Abbas, S. Grace, J. Graef, S. Negi, Oscillation of second-order non-canonical non-linear dynamic equations with a sub-linear neutral term, Differ. Equ. Dyn. Syst., 32 (2024), 819–829. http://dx.doi.org/10.1007/s12591-022-00592-0 doi: 10.1007/s12591-022-00592-0
    [2] S. Affan, T. Hassan, E. Elabbasy, E. Saied, A. Hassan, Oscillatory behavior of second-order nonlinear delay dynamic equations with multiple sublinear neutral terms utilizing canonical transformation, Math. Meth. Appl. Sci., 48 (2025), 1589–1600. http://dx.doi.org/10.1002/mma.10397 doi: 10.1002/mma.10397
    [3] R. Agarwal, S. Grace, D. O'Regan, Oscillation theory for second order dynamic equations, CRC Press, 2003. https://doi.org/10.4324/9780203222898
    [4] B. Baculíková, Oscillation of second-order nonlinear noncanonical differential equations with deviating argument, Appl. Math. Lett., 91 (2019), 68–75. http://dx.doi.org/10.1016/j.aml.2018.11.021 doi: 10.1016/j.aml.2018.11.021
    [5] B. Baculíková, Oscillatory behavior of the second order noncanonical differential equations, Electron. J. Qual. Theory Differ. Equ., 89 (2019), 1–11. http://dx.doi.org/10.14232/ejqtde.2019.1.89 doi: 10.14232/ejqtde.2019.1.89
    [6] B. Baculíková, J. Dzurina, Oscillatory criteria via linearization of half-linear second order delay differential equations, Opuscula Math., 40 (2020), 523–536. https://doi.org/10.7494/OpMath.2020.40.5.523 doi: 10.7494/OpMath.2020.40.5.523
    [7] M. Bohner, S. Grace, I. Jadlovská, Sharp oscillation criteria for second-order neutral delay differential equations, Math. Meth. Appl. Sci., 43 (2020), 10041–10053. http://dx.doi.org/10.1002/mma.6677 doi: 10.1002/mma.6677
    [8] M. Bohner, A. Peterson, Dynamic equations on time scales: An introduction with applications, Springer Science & Business Media, 2001. https://doi.org/10.1007/978-1-4612-0201-1
    [9] M. Bohner, A. Peterson, Advances in dynamic equations on time scales, Springer Science & Business Media, 2002. https://doi.org/10.1007/978-0-8176-8230-9
    [10] Z. Cai, L. Huang, Z. Wang, X. Pan, S. Liu, Periodicity and multi-periodicity generated by impulses control in delayed cohen–grossberg-type neural networks with discontinuous activations, Neural Netw., 143 (2021), 230–245. http://dx.doi.org/10.1016/j.neunet.2021.06.013 doi: 10.1016/j.neunet.2021.06.013
    [11] G. Chatzarakis, S. Grace, I. Jadlovská, On the sharp oscillation criteria for half-linear second-order differential equations with several delay arguments, Appl. Math. Comput., 397 (2021), 125915. http://dx.doi.org/10.1016/j.amc.2020.125915 doi: 10.1016/j.amc.2020.125915
    [12] G. Chatzarakis, S. Grace, I. Jadlovská, A sharp oscillation criterion for second-order half-linear advanced differential equations, Acta Math. Hungar., 163 (2021), 552–562. http://dx.doi.org/10.1007/s10474-020-01110-w doi: 10.1007/s10474-020-01110-w
    [13] G. Chatzarakis, N. Indrajith, S. Panetsos, E. Thandapani, Improved oscillation criteria of second-order advanced non-canonical difference equations, Aust. J. Math. Anal. Appl., 19 (2022), 5.
    [14] G. Chatzarakis, N. Indrajith, E. Thandapani, K. Vidhyaa, Oscillatory behavior of second-order non-canonical retarded difference equations, Aust. J. Math. Anal. Appl., 18 (2021), 20.
    [15] P. Gopalakrishnan, A. Murugesan, C. Jayakumar, Oscillation conditions of the second order noncanonical difference equations, J. Math. Comput. SCI-JM., 25 (2022), 351–360. http://dx.doi.org/10.22436/jmcs.025.04.05 doi: 10.22436/jmcs.025.04.05
    [16] A. Hassan, S. Affan, New oscillation criteria and some refinements for second-order neutral delay dynamic equations on time scales, J. Math. Comput. SCI-JM., 28 (2023), 192–202. http://dx.doi.org/10.22436/jmcs.028.02.07 doi: 10.22436/jmcs.028.02.07
    [17] A. Hassan, S. Affan, Oscillation criteria for second-order delay dynamic equations with a sub-linear neutral term on time scales, Filomat, 37 (2023), 7445–7454. http://dx.doi.org/10.2298/FIL2322445H doi: 10.2298/FIL2322445H
    [18] A. Hassan, O. Moaaz, S. Askar, A. Alshamrani, S. Affan, Enhanced oscillation criteria for non-canonical second-order advanced dynamic equations on time scales, Symmetry, 16 (2024), 1457. http://dx.doi.org/10.3390/sym16111457 doi: 10.3390/sym16111457
    [19] A. Hassan, I. Odinaev, T. Hassan, Oscillatory behavior of noncanonical quasilinear second-order dynamic equations on time scales, J. Math., 2023 (2023), 5585174. http://dx.doi.org/10.1155/2023/5585174 doi: 10.1155/2023/5585174
    [20] A. Hassan, H. Ramos, O. Moaaz, Second-order dynamic equations with noncanonical operator: Oscillatory behavior, Fractal Fract., 7 (2023), 134. http://dx.doi.org/10.3390/fractalfract7020134 doi: 10.3390/fractalfract7020134
    [21] T. Hassan, M. Bohner, I. Florentina, A. Abdel Menaem, M. Mesmouli, New criteria of oscillation for linear sturm–liouville delay noncanonical dynamic equations, Mathematics, 11 (2023), 4850. http://dx.doi.org/10.3390/math11234850 doi: 10.3390/math11234850
    [22] T. Hassan, C. Cesarano, M. Mesmouli, H. Zaidi, I. Odinaev, Iterative hille-type oscillation criteria of half-linear advanced dynamic equations of second order, Math. Meth. Appl. Sci., 74 (2024), 5651–5663. http://dx.doi.org/10.1002/mma.9883 doi: 10.1002/mma.9883
    [23] T. Hassan, R. El-Nabulsi, N. Iqbal, A. Abdel Menaem, New criteria for oscillation of advanced noncanonical nonlinear dynamic equations, Mathematics, 12 (2024), 824. http://dx.doi.org/10.3390/math12060824 doi: 10.3390/math12060824
    [24] S. Hilger, Analysis on measure chains — a unified approach to continuous and discrete calculus, Results Math., 18 (1990), 18–56. https://doi.org/10.1007/BF03323153 doi: 10.1007/BF03323153
    [25] N. Indrajith, J. Graef, E. Thandapani, Kneser-type oscillation criteria for second-order half-linear advanced difference equations, Opuscula Math., 42 (2022), 55–64. http://dx.doi.org/10.7494/OpMath.2022.42.1.55 doi: 10.7494/OpMath.2022.42.1.55
    [26] I. Jadlovská, Oscillation criteria of kneser-type for second-order half-linear advanced differential equations, Appl. Math. Lett., 106 (2020), 106354. http://dx.doi.org/10.1016/j.aml.2020.106354 doi: 10.1016/j.aml.2020.106354
    [27] I. Jadlovská, New criteria for sharp oscillation of second-order neutral delay differential equations, Mathematics, 9 (2021), 2089. http://dx.doi.org/10.3390/math9172089 doi: 10.3390/math9172089
    [28] C. Jayakumar, A. Murugesan, Oscillation result for half-linear delay difference equations of second-order, Malaya J. Mat., 9 (2021), 1153–1159.
    [29] C. Jayakumar, S. Santra, D. Baleanu, R. Edwan, V. Govindan, A. Murugesan, et al., Oscillation result for half-linear delay difference equations of second-order, Math. Biosci. Eng., 19 (2022), 3879–3891. http://dx.doi.org/10.3934/mbe.2022178 doi: 10.3934/mbe.2022178
    [30] M. Liu, H. Shi, Exponential stability of dynamical systems on time scales with application to multi-agent systems, Axioms, 13 (2024), 100. http://dx.doi.org/10.3390/axioms13020100 doi: 10.3390/axioms13020100
    [31] J. Shi, S. Gu, S. Xing, C. Chen, Dynamic event-triggered fault detection for multi time scale systems: Application to grid connected converters, J. Frank. Inst., 361 (2024), 106738. http://dx.doi.org/10.1016/j.jfranklin.2024.106738 doi: 10.1016/j.jfranklin.2024.106738
    [32] S. Shi, Z. Han, A new approach to the oscillation for the difference equations with several variable advanced arguments, J. Appl. Math. Comput., 68 (2022), 2083–2096. http://dx.doi.org/10.1007/s12190-021-01605-x doi: 10.1007/s12190-021-01605-x
    [33] A. Soliman, A. Hassan, S. Affan, Oscillatory behavior of second order delay dynamic equa-tions with a sub-linear neutral term on time scales, J. Math. Comput. SCI-JM., 24 (2022), 97–109. http://dx.doi.org/10.22436/jmcs.024.02.01 doi: 10.22436/jmcs.024.02.01
    [34] Y. Tian, X. Su, C. Shen, X. Ma, Exponentially extended dissipativity-based filtering of switched neural networks, Automatica, 161 (2024), 111465. http://dx.doi.org/10.1016/j.automatica.2023.111465 doi: 10.1016/j.automatica.2023.111465
    [35] H. Zhao, J. Zhang, J. Li, Decay estimates of solution to the two-dimensional fractional quasi-geostrophic equation, Math. Meth. Appl. Sci., 47 (2024), 4043–4057. http://dx.doi.org/10.1002/mma.9802 doi: 10.1002/mma.9802
    [36] K. Zhao, Existence and stability of a nonlinear distributed delayed periodic ag-ecosystem with competition on time scales, Axioms, 12 (2023), 315. https://doi.org/10.3390/axioms12030315 doi: 10.3390/axioms12030315
    [37] K. Zhao, Asymptotic stability of a periodic ga-predation system with infinite distributed lags on time scales, Int. J. Control, 97 (2024), 1542–1552. http://dx.doi.org/10.1080/00207179.2023.2214251 doi: 10.1080/00207179.2023.2214251
  • This article has been cited by:

    1. Traian A. Pirvu, Shuming Zhang, Spread Option Pricing Under Finite Liquidity Framework, 2024, 12, 2227-9091, 173, 10.3390/risks12110173
  • Reader Comments
  • © 2025 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(440) PDF downloads(28) Cited by(0)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog