Research article

Pricing and hedging bond options and sinking-fund bonds under the CIR model

  • Received: 27 November 2021 Revised: 13 December 2021 Accepted: 04 January 2022 Published: 10 January 2022
  • JEL Codes: G12, G13

  • This article derives simple closed-form solutions for computing Greeks of zero-coupon and coupon-bearing bond options under the CIR interest rate model, which are shown to be accurate, easy to implement, and computationally highly efficient. These novel analytical solutions allow us to extend the literature in two other directions. First, the static hedging portfolio approach is used for pricing and hedging American-style plain-vanilla zero-coupon bond options under the CIR model. Second, we derive analytically the comparative static properties of sinking-fund bonds under the same interest rate modeling setup.

    Citation: Manuela Larguinho, José Carlos Dias, Carlos A. Braumann. Pricing and hedging bond options and sinking-fund bonds under the CIR model[J]. Quantitative Finance and Economics, 2022, 6(1): 1-34. doi: 10.3934/QFE.2022001

    Related Papers:

    [1] 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
    [2] Joseph Atta-Mensah . Commodity-linked bonds as an innovative financing instrument for African countries to build back better. Quantitative Finance and Economics, 2021, 5(3): 516-541. doi: 10.3934/QFE.2021023
    [3] Hammad Siddiqi . Financial market disruption and investor awareness: the case of implied volatility skew. Quantitative Finance and Economics, 2022, 6(3): 505-517. doi: 10.3934/QFE.2022021
    [4] Longfei Wei, Lu Liu, Jialong Hou . Pricing hybrid-triggered catastrophe bonds based on copula-EVT model. Quantitative Finance and Economics, 2022, 6(2): 223-243. doi: 10.3934/QFE.2022010
    [5] 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
    [6] Richard J. Cebula, Don Capener . The impact of federal income tax rate cuts on the municipal bond market in the U.S.: A brief exploratory empirical note. Quantitative Finance and Economics, 2018, 2(2): 407-412. doi: 10.3934/QFE.2018.2.407
    [7] Lianzhang Bao, Guangliang Zhao, Zhuo Jin . A new equilibrium trading model with asymmetric information. Quantitative Finance and Economics, 2018, 2(1): 217-229. doi: 10.3934/QFE.2018.1.217
    [8] Jin Liang, Hong-Ming Yin, Xinfu Chen, Yuan Wu . On a Corporate Bond Pricing Model with Credit Rating Migration Risksand Stochastic Interest Rate. Quantitative Finance and Economics, 2017, 1(3): 300-319. doi: 10.3934/QFE.2017.3.300
    [9] Ick Jin . Systematic ESG risk and hedge fund. Quantitative Finance and Economics, 2024, 8(2): 387-409. doi: 10.3934/QFE.2024015
    [10] 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
  • This article derives simple closed-form solutions for computing Greeks of zero-coupon and coupon-bearing bond options under the CIR interest rate model, which are shown to be accurate, easy to implement, and computationally highly efficient. These novel analytical solutions allow us to extend the literature in two other directions. First, the static hedging portfolio approach is used for pricing and hedging American-style plain-vanilla zero-coupon bond options under the CIR model. Second, we derive analytically the comparative static properties of sinking-fund bonds under the same interest rate modeling setup.



    A bond is a contract which pays its holder a known amount, the principal, at a known future date, called the maturity of the contract. The bond may also pay periodically to its holder fixed cash dividends, called the coupons. This type of bonds are known as coupon bonds (sometimes also called coupon-paying or coupon-bearing bonds). If the bond pays no coupons, it is known as a zero-coupon bond (or pure discount bond). Several bonds may contain special clauses or some embedded options. There are also some derivative contracts whose underlying asset is a bond.

    This paper offers three contributions to the existent bond option pricing literature. First, we provide closed-form solutions to efficiently and accurately compute sensitivity measures (commonly known as Greeks) of pure discount and coupon-bearing bond options under the Cox et al. (1985) mean-reverting square-root model (hereafter, CIR model), which, to the best of our knowledge, are new in the option pricing literature.1 Following the insights of Larguinho et al. (2013) and Dias et al. (2020), the obtained closed-form expressions for Greeks under the CIR model are expressed in terms of noncentral chi-square distribution functions, which can be efficiently computed via Benton and Krishnamoorthy (2003, Algorithm 7.3).2

    1 To simplify the notation and better emphasize our exposition, we focus on the classical constant coefficient CIR model to derive Greeks, but the same line of reasoning can be applied also under the time-varying coefficients version of the CIR model offered by Jamshidian (1995) and Maghsoodi (1996), as well as under the CIR++ model of Brigo and Mercurio (2001).

    2 This algorithm has been used in several option pricing applications involving such distribution, e.g., Ruas et al. (2013), Dias et al. (2015), Nunes et al. (2015), Cruz and Dias (2017), Cruz and Dias (2020), and Dias et al. (2020).

    These novel Greeks should be valuable to both academics and practitioners. For instance, a dealer of the financial industry should be able not only to price a given option contract but also to hedge it. Knowing and understanding such sensitivity measures is thus pivotal in the design of hedging strategies for a given security or a portfolio of securities, when closing the position is not viable or desirable. Greeks also enjoy many other multiple applications such as market risk measurement, profit and loss attribution, model risk assessment and optimal contract design, and to determine parameter values from market prices. Moreover, the availability of analytical solutions for Greeks reduces substantially the computational burden when dealing with large portfolios of securities that have to be re-evaluated frequently and allows them to be easily coded in any desired computer language.

    Armed with these new analytical solutions for Greeks, we can now extend the literature in two other directions.3 Hence, and as our second contribution, we are able to price (and hedge) American-style option contracts on zero-coupon bonds under the CIR model via the static hedge portfolio (hereafter, SHP) approach offered by Chung and Shih (2009) and Ruas et al. (2013) in the context of stock options. It is well-know that the pricing (and hedging) of American-style contingent claims boils down to a boundary value problem in a domain whose boundary is not fully known and, therefore, must be also determined. In other words, the option price and the early exercise boundary must be determined simultaneously as the solution of the same free boundary problem that has been set up by McKean (1965). As for the stock options case, there are no closed-form solutions for pricing American-style options on bonds. Hence, these contracts have been usually evaluated numerically using finite difference, finite volume, and finite element methods—see, for instance, Hull and White (1990), Allegretto et al. (2003), Yang (2004), ShuJin and ShengHong (2006), Zhou et al. (2011), and Thakoor et al. (2012)—, through a binomial or trinomial tree approach—see, for example, Nelson and Ramaswamy (1990), Tian (1992), Tian (1994), and Nawalkha and Beliaeva (2007)—, via the least-squares Monte Carlo scheme of Longstaff and Schwartz (2001), or with the optimal stopping approach proposed by Chesney et al. (1993). More recently, Deng (2015) considers the valuation of American-style put options on zero-coupon bonds in a jump-extended CIR model, Najafi et al. (2018) evaluate the American-style put option on a zero-coupon bond assuming that the interest rate model is governed by a fractional CIR process, whereas Peng and Schellhorn (2018) study the probability distribution of the interest rate under an extended CIR model with time-varying dimension and propose a pricing method for options on zero-coupon bonds.

    3 Another possible research direction, outside the scope of the present article, is the comparison of alternative binomial approximation schemes for computing the option hedge ratios in the spirit of Pelsser and Vorst (1994), Chung and Shackleton (2002), Chung and Shackleton (2005), Chung et al. (2011), and Cruz and Dias (2017).

    Alternatively, we show how to tackle the valuation of American-style options on pure discount bonds with a distinct approach that has proved to be extremely efficient and accurate in the case of American-style stock options and under different assumptions for the dynamics of the underlying asset price—see Chung and Shih (2009) and Ruas et al. (2013) for more details.4 Broadly speaking, we use standard European-style zero-coupon bond options with multiple strikes and multiple maturities, because the optimal exercise boundary of such American-style contracts are not known ex-ante. This approach creates a static portfolio of European-style options whose values match the payoff of the American-style option being hedged at expiration and along the boundary, by applying the value-matching and smooth-pasting conditions on the early exercise boundary. As for the case of stock options, we show that the SHP methodology is also robust and computationally efficient when dealing with bond options.5

    4 The SHP approach has been shown to be useful also for pricing and hedging barrier option contracts, as highlighted in Chung et al. (2010), Chung et al. (2013), Dias et al. (2015), Nunes et al. (2015), Guo and Chang (2020), and Nunes et al. (2020).

    5 Unfortunately, the existence of interim coupons prevents the implementation of the SHP approach for American-style options on coupon-bearing bonds. Nevertheless, the SHP methodology should be useful both in theory and in practice since it provides a fast and accurate method for pricing American-style options on zero-coupon bonds, thus being a viable alternative to the aforementioned schemes available in the literature for these contracts. Moreover, this methodology can also be applied to any other single-factor interest rate model offering closed-form solutions for option prices and hedge ratios.

    As our final contribution, we revisit the Bacinello et al. (1996) work and provide analytic tractable formulae for valuing and analyzing comparative statics of sinking-fund bonds in the CIR framework. We shall note that while Bacinello et al. (1996) have been able to study such issues in closed-form under the Vasicek (1977) model, they analyze numerically the comparative static properties of the sinking-fund bond in the CIR model given the absence of closed-form expressions of Greeks under the CIR modeling setup. Using our novel solutions, we show that the stochastic duration of the sinking-fund bond is between the stochastic duration of the corresponding serial and coupon bonds. Although this issue has been shown already by Bacinello et al. (1996) through numerical differentiation, we are able to establish this property analytically using the proposed closed-form solutions for the CIR sensitivity measures.

    The remainder of the paper is organized as follows. Section 2 outlines a brief summary of the CIR interest rate dynamics and the analytical formulae for computing discount bonds, coupon-bearing bonds, and European-style options on discount bonds and coupon-paying bonds in a CIR economy. Section 3 derives analytical tractable solutions of the sensitivity measures of bond options under the same interest rate dynamics setting and presents some numerical examples to enhance the efficiency of our closed-form solutions. Section 4 implements the SHP approach for pricing and hedging American-style options on zero-coupon bonds. Section 5 provides analytically tractable formulae to analyze the comparative statics properties of a sinking-fund bond in the CIR framework. Section 6 presents the concluding remarks. All accessory results are relegated to the Appendix.

    This section presents a brief remainder of the analytical formulae for computing discount bonds, coupon-bearing bonds, and European-style call and put options on zero-coupon bonds and coupon-paying bonds in a CIR economy that will be required later. Even though these results are well known in the literature they are needed to establish notations and the desire for self-consistency.

    Hereafter, we consider a CIR economy in which EQt denotes the time-t expectation under the martingale (or risk-neutral) probability measure Q, with respect to the risk-adjusted process for the instantaneous interest rate rt

    drt=κ(θrt)dt+σrtdWQt, (1)

    where κ:=κ+λ is the risk-neutral parameter that determines the speed of adjustment (reversion rate or reverting rate), θ:=κθ/(κ+λ) is the risk-neutral long-run mean of the instantaneous interest rate (asymptotic interest rate or reverting level), σ is the volatility of the process, λ is the market price of risk parameter, and WQt is a standard Brownian motion under Q.6 It is well known that the κθ term plays a key role under this diffusion and has important implications for capture of the interest rate process r at a value of zero. The condition 2κθσ2 ensures that the interest rate remains positive.7

    6 We recall that, when λ=0, we have κ=κ and θ=θ, which implies that the speed of adjustment and the asymptotic interest rate under the physical and risk-neutral measures are the same.

    7 See Feller (1951) for a complete description of the boundary conditions.

    In a CIR economy, the time-t price of a zero-coupon bond maturing at time s (with s>t), Z(r,t,s), is given by

    Z(r,t,s)=EQt[estrudu]=A(t,s)eB(t,s)r, (2)

    with r=rt at the valuation date t, and where the constants A(t,s), B(t,s), and γ>0 are given by

    A(t,s):=[2γe[(κ+λ+γ)(st)]/2(κ+λ+γ)(eγ(st)1)+2γ]2κθ/σ2, (3)
    B(t,s):=2(eγ(st)1)(κ+λ+γ)(eγ(st)1)+2γ, (4)

    and

    γ:=[(κ+λ)2+2σ2]1/2. (5)

    It is well established in the literature—see, for instance, Jamshidian (1989)—that, for all one-factor term structure models, a coupon-paying bond can be decomposed into a portfolio of zero-coupon bonds of different maturities. Hence, the time-t value of a coupon-bearing bond expiring at time s (with s>t), P(r,t,s), can be simply expressed as a weighted sum of zero-coupon bond prices, that is

    P(r,t,s)=Ni=1aiZ(r,t,si), (6)

    with s1,s2,,sN (and sN=s) representing the N dates on which payments are made, and each ai>0 term denoting the amount of the payments made.8

    8 For example, consider a 10-year 6% coupon bond with a par value of 100 and semiannual coupon payments. In this case, N=20 since the bond makes 19 semiannual coupon payments of 3 as well as a final payment of 103. Thus, ai=100×6%/2=3 for i=1,2,,19, a20=3+100=103, and s1=0.5,s2=1,,s19=9.5, and s20=10.

    Analytic solutions for pricing options on discount bonds have been proposed by Cox et al. (1985). Denote vzc(r,t,T,s,K;α) as the time-t price of a European-style call option (if α=1) or put option (if α=1) with strike price K, expiration date T, written on a zero-coupon bond with maturity date s (with s>T>t), and with the instantaneous interest rate at time t given by r.9 The time-t price of a zero-coupon bond option is given by

    9 It is well-known that K is restricted to be less than A(T,s), the maximum possible bond price at time T, since otherwise the option would never be exercised and would be worthless—see Cox et al. (1985, Page 396).

    vzc(r,t,T,s,K;α)=αZ(r,t,s)Q[x1(t,T,s,K);a,b1(r,t,T,s);α]αKZ(r,t,T)Q[x2(t,T,s,K);a,b2(r,t,T);α], (7)

    where Q(x;a,b;α) is the distribution function (for α=1) and the complementary distribution function (for α=1) of the noncentral chi-square distribution with a degrees of freedom and non-centrality parameter b,

    x1(t,T,s,K):=2r(T,s,K)[ϕ(t,T)+ψ+B(T,s)], (8)
    x2(t,T,s,K):=2r(T,s,K)[ϕ(t,T)+ψ], (9)
    a:=4κθσ2, (10)
    b1(r,t,T,s):=2ϕ2(t,T)reγ(Tt)ϕ(t,T)+ψ+B(T,s), (11)
    b2(r,t,T):=2ϕ2(t,T)reγ(Tt)ϕ(t,T)+ψ, (12)
    ϕ(t,T):=2γσ2(eγ(Tt)1), (13)
    ψ:=κ+λ+γσ2, (14)

    and

    r(T,s,K):=1B(T,s)ln(A(T,s)K), (15)

    with r being the critical interest rate below which exercise will occur, i.e., K=Z(r,T,s).

    Following the argument of Jamshidian (1989) that an option on a portfolio of zero-coupon bonds decomposes into a portfolio of options on the individual discount bonds in the portfolio, then the time-t price of a European-style call option (if α=1) or put option (if α=1), with strike price K and maturity date T, on a portfolio consisting of N zero-coupon bonds with different expiry dates si, is given by

    vcb(r,t,T,s,K;α)=Ni=1aivzc(r,t,T,si,Ki;α), (16)

    with T<s1<s2<<sN=s, ai>0, Ki=Z(r,T,si), and where r is the solution to Ni=1aiZ(r,T,si)=K.10

    10 Alternatively, we may use the equivalent closed-form expressions offered by Longstaff (1993, Equations 7 and 9).

    Remark 1. Note that the underlying asset for coupon bond options is actually the portfolio of discount bonds expiring after the option's maturity date. However, the value of this portfolio is strictly less than the current price of the coupon bond if the bond pays coupons before the expiry date of the option. As argued by Longstaff (1993, Page 32), the value of the underlying asset for a 5-year option on a 10-year bond is not the current price of a 15-year bond, but the price of a 15-year bond minus the present value of coupon payments to be made during the next 5 years. In other words, the option's payoff—and, hence, the coupon bond option price—does not depend on the payments of the coupon bond to be made before the expiry date of the option.

    This section derives closed-form solutions for Greeks under the CIR model, which, to the best of our knowledge, are new in the literature.

    Let us begin with two important general relations, which will be used for deriving Greeks under the CIR model. Following Johnson et al. (1995, pp. 442-443) or Larguinho et al. (2013, Equations A2a and A2b), we know that

    Q[x;a,b;α]x=αp(x;a,b), (17)

    and

    Q[x;a,b;α]b=αp(x;a+2,b), (18)

    where p(x;a,b) is the probability density function of a noncentral chi-square distribution as given by Johnson et al. (1995, Equation 29.4), that is

    p(x;a,b)=12e(b+x)/2(xb)(a2)/4I(a2)/2(bx),x>0, (19)

    with Iq() being the modified Bessel function of the first kind of order q, as defined in Abramowitz and Stegun (1972, Equation 9.6.10). We will also need to use the first derivative of the probability density function (19) with respect to the non-centrality parameter b, which can be computed through the following recurrence relation given by Cohen (1988):

    p(x;a,b)b=12[p(x;a,b)+p(x;a+2,b)]. (20)

    Next propositions and remarks offer the proposed novel closed-form solutions for computing sensitivity measures of zero-coupon bond options under the CIR model, namely the rho (or interest rate delta), interest rate gamma, theta, and eta (or strike delta).11 We notice that the corresponding rho, interest rate gamma, and theta measures of coupon-bearing bond options arise immediately, because it is possible to apply the decomposition technique of Jamshidian (1989) to these Greeks. For the case of eta, however, it is necessary to combine the decomposition technique with the classic chain rule.

    11 Note that the so-called vega—which is the sensitivity of the bond option price with respect to the volatility parameter σ—depends on the degrees of freedom parameter a of the noncentral chi-square distribution function, for which (to the authors knowledge) there is no simple relationship as those given in equations (17) and (18). See Alvarez (2001) who discusses the conditions which determine the sign of the effect of increased volatility on the price of a general interest rate claim under a broad class of interest rate models.

    The rho or interest rate delta can be computed as:

    Proposition 1. Consider the pricing solution of a zero-coupon bond option under the CIR model as defined in equation (7). Then, the rho (or interest rate delta) of a zero-coupon bond call option (if α=1) or put option (if α=1) is given by

    ρzcv(.):=vzc(.)r=Z(r,t,s)[αB(t,s)Q[x1(.);a,b1(.);α]b1(.)rp(x1(.);a+2,b1(.))]KZ(r,t,T)[αB(t,T)Q[x2(.);a,b2(.);α]b2(.)rp(x2(.);a+2,b2(.))]. (21)

    Proof. Please see Appendix A.

    Remark 2. The rho of a coupon bond call option (if α=1) or put option (if α=1) arises immediately if one applies the decomposition technique of Jamshidian (1989), that is

    ρcbv(.):=vcb(.)r=Ni=1aivzc(r,t,T,si,Ki;α)r=Ni=1aiρzcv(r,t,T,si,Ki;α). (22)

    The previous remark shows that it is straightforward to compute call and put interest rate deltas in closed-form for coupon-paying bond options under the CIR framework. This will allow us to compare the results calculated by expression (22) with the rho values shown in Wei (1997, Table Ⅱ), which have been obtained through a numerical integration scheme.

    The interest rate gamma can be computed as:

    Proposition 2. Consider the pricing solution of a zero-coupon bond option under the CIR model as defined in equation (7). Then, the interest rate gamma of a zero-coupon bond call option (if α=1) or put option (if α=1) is given by

    Γzcv,r(.):=2vzc(.)r2=Z(r,t,s)[αB2(t,s)Q[x1(.);a,b1(.);α]+2B(t,s)b1(.)rp(x1(.);a+2,b1(.))12(b1(.)r)2(p(x1(.);a+2,b1(.))+p(x1(.);a+4,b1(.)))]KZ(r,t,T)[αB2(t,T)Q[x2(.);a,b2(.);α]+2B(t,T)b2(.)rp(x2(.);a+2,b2(.))12(b2(.)r)2(p(x2(.);a+2,b2(.))+p(x2(.);a+4,b2(.)))]. (23)

    Proof. Please see Appendix B.

    Remark 3. The interest rate gamma of a coupon bond call option (if α=1) or put option (if α=1) arises immediately if one applies the decomposition technique of Jamshidian (1989), that is

    Γcbv,r(.):=2vcb(.)r2=Ni=1aiρzcv(r,t,T,si,Ki;α)r=Ni=1aiΓzcv,r(r,t,T,si,Ki;α). (24)

    The theta can be computed as:

    Proposition 3. Consider the pricing solution of a zero-coupon bond option under the CIR model as defined in equation (7). Then, the theta of a zero-coupon bond call option (if α=1) or put option (if α=1) is given by

    θzcv(.):=vzc(.)t=Z(r,t,s)[αζsQ[x1(.);a,b1(.);α]+ξp(x1(.);a,b1(.))ϱ1p(x1(.);a+2,b1(.))]KZ(r,t,T)[αζTQ[x2(.);a,b2(.);α]+ξp(x2(.),a,b2(.))ϱ2p(x2(.);a+2,b2(.))], (25)

    where

    ζj=κθσ2(κ+λ+γ)(eγ(jt)1)(2γ(κ+λ+γ))(κ+λ+γ)(eγ(jt)1)+2γ+4rγ2eγ(jt)[(κ+λ+γ)(eγ(jt)1)+2γ]2, (26)

    for j{T,s},

    ξ=4rγ2eγ(Tt)σ2(eγ(Tt)1)2, (27)
    ϱ1=b1(r,t,T,s)γ(ϕ(t,T)+ψ+B(T,s))+(ψ+B(T,s))eγ(Tt)(eγ(Tt)1)(ϕ(t,T)+ψ+B(T,s)), (28)

    and

    ϱ2=b2(r,t,T)γ(ϕ(t,T)+ψ)+ψeγ(Tt)(eγ(Tt)1)(ϕ(t,T)+ψ). (29)

    Proof. Please see Appendix C.

    Remark 4. The theta of a coupon bond call option (if α=1) or put option (if α=1) arises immediately if one applies the decomposition technique of Jamshidian (1989), that is

    θcbv(.):=vcb(.)t=Ni=1aivzc(r,t,T,si,Ki;α)t=Ni=1aiθzcv(r,t,T,si,Ki;α). (30)

    The eta can be computed as:

    Proposition 4. Consider the pricing solution of a zero-coupon bond option under the CIR model as defined in equation (7). Then, the eta of a zero-coupon bond call option (if α=1) or put option (if α=1) is given by

    ηzcv(.):=vzc(.)K=2Z(r,t,s)p(x1(.);a,b1(.))ϕ(t,T)+ψ+B(T,s)B(T,s)KZ(r,t,T)[αQ[(x2(.);a,b2(.);α]2p(x2(.);a,b2(.))ϕ(t,T)+ψB(T,s)]. (31)

    Proof. Please see Appendix D.

    Proposition 5. The eta of a coupon bond call option (if α=1) or put option (if α=1) arises immediately if one combines the decomposition technique of Jamshidian (1989) and the classic chain rule, obtaining

    ηcbv(.):=vcb(.)K=Ni=1aivzc(r,t,T,si,Ki;α)K=Ni=1aiB(T,si)Z(r,T,si)ηzcv(r,t,T,si,Ki;α)Nj=1ajB(T,sj)Z(r,T,sj). (32)

    Proof. Please see Appendix E.

    We recall that in a CIR economy it is the independent variable r that is assumed to be stochastic. However, the underlying asset of a bond option contract is the bond price itself and not the interest rate. Nevertheless, it is still possible to compute the delta (or hedge ratio) of the bond option with respect to the underlying bond price. This is accomplished because we can apply the classic chain rule for deriving the delta of a zero-coupon bond option. Moreover, and even though it is not possible to apply the decomposition technique of Jamshidian (1989) when computing the delta of a coupon-paying bond option, we can still determine its value by simply using, again, the classic chain rule.

    Next remark shows the analytical solutions for computing the delta of zero-coupon and coupon-paying bond options under the CIR model.12

    12 Additional details about the computation of the delta of a zero-coupon bond option will be discussed later in Remark 6.

    Remark 5. The delta (with respect to the underlying bond price) for both zero-coupon and coupon-paying bond options arise immediately if one uses the results obtained in Proposition 1 and Remark 2, that is

    Δzcv(.):=vzc(r,t,T,s,K;α)Z(r,t,s)=vzc(r,t,T,s,K;α)rrZ(r,t,s)=ρzcv(r,t,T,s,K;α)1Z(r,t,s)r=ρzcv(r,t,T,s,K;α)B(t,s)Z(r,t,s), (33)

    and

    Δcbv(.):=vcb(r,t,T,s,K;α)P(r,t,s)=vcb(r,t,T,s,K;α)rrP(r,t,s)=ρcbv(r,t,T,s,K;α)1P(r,t,s)r=ρcbv(r,t,T,s,K;α)Ni=1aiB(t,si)Z(r,t,si). (34)

    Interestingly, the gamma (with respect to the underlying bond price) for both zero-coupon and coupon-paying bond options can be expressed analytically in terms of others sensitivity measures, as shown in the following two propositions.

    Proposition 6. The gamma (with respect to the underlying bond price) of a zero-coupon call option (if α=1) or put option (if α=1) can be computed as

    Γzcv,Z(.):=Δzcv(.)Z(r,t,s)=Γzcv,r(.)[B(t,s)Z(r,t,s)]2Δzcv(.)Z(r,t,s). (35)

    Proof. Please see Appendix F.

    Proposition 7. The gamma (with respect to the underlying bond price) of a coupon-paying call option (if α=1) or put option (if α=1) can be computed as

    Γcbv,Z(.):=Δcbv(.)P(r,t,s)=Γcbv,r(.)Δcbv(.)Ni=1ai[B(t,si)]2Z(r,t,si)[Ni=1aiB(t,si)Z(r,t,si)]2. (36)

    Proof. Please see Appendix G.

    This subsection presents some numerical results of the novel closed-form solutions of Greeks of bond options under the CIR model. For completeness, we note that our software programs were implemented in Matlab R2021b and run on a personal computer with an Intel Core i9-10900 2.80 GHz processor and 64 GB of ram memory.

    Table 1 values 4-year call option prices (if α=1) and put option prices (if α=1), as well as their corresponding rho, interest rate gamma, theta, eta, delta, and gamma sensitivity measures, written on a 10-year zero-coupon bond with face value equal to $1.0 for different levels of the interest rate (r), strike price K = $0.6 , and using the parameter values κ=0.2339, θ=0.0808, σ=0.0854, and λ=0, borrowed from Chan et al. (1992, Table Ⅲ). We should mention that the obtained zero-coupon bond prices and the option prices are both expressed as percentages of the face value. For completeness, we shall also mention that the required noncentral chi-square distribution function and the corresponding probability density function have been computed using, respectively, the Benton and Krishnamoorthy (2003) algorithm and the built-in function ncx2pdf available in Matlab.

    Table 1.  Prices and Greeks of European-style options on zero-coupon bonds under the CIR model.
    Call options
    r Z(r,t,s) vzc(.;1) ρzcv(.;1) Γzcv,r(.;1) θzcv(.;1) ηzcv(.;1) Δzcv(.;1) Γzcv,Z(.;1) pde test
    0.01 59.3183 7.2123 0.7992 6.3552 0.0137 0.8185 0.3624 0.6957 1.95E17
    0.02 57.1534 6.4447 0.7364 6.1884 0.0113 0.7765 0.3466 0.7642 3.92E17
    0.03 55.0675 5.7389 0.6755 5.9827 0.0091 0.7327 0.3299 0.8281 8.67E18
    0.04 53.0577 5.0929 0.6169 5.7418 0.0071 0.6878 0.3127 0.8861 9.11E18
    0.05 51.1213 4.5043 0.5608 5.4707 0.0053 0.6422 0.2951 0.9372 2.43E17
    0.06 49.2555 3.9703 0.5075 5.1752 0.0037 0.5965 0.2772 0.9805 9.97E18
    0.07 47.4578 3.4881 0.4574 4.8617 0.0024 0.5512 0.2592 1.0154 1.73E17
    0.08 45.7258 3.0546 0.4104 4.5365 0.0012 0.5067 0.2414 1.0417 1.78E17
    0.09 44.0569 2.6663 0.3666 4.2056 0.0002 0.4634 0.2238 1.0594 2.26E17
    0.10 42.4490 2.3202 0.3262 3.8744 0.0006 0.4218 0.2067 1.0685 3.04E18
    0.11 40.8997 2.0128 0.2891 3.5480 0.0012 0.3821 0.1901 1.0695 2.99E17
    0.12 39.4070 1.7408 0.2553 3.2303 0.0017 0.3445 0.1742 1.0627 5.55E17
    0.13 37.9688 1.5012 0.2245 2.9249 0.0020 0.3093 0.1590 1.0489 2.52E17
    0.14 36.5830 1.2909 0.1967 2.6343 0.0023 0.2764 0.1446 1.0287 3.27E17
    0.15 35.2479 1.1069 0.1717 2.3607 0.0024 0.2460 0.1311 1.0027 2.08E17
    Put options
    r Z(r,t,s) vzc(.;1) ρzcv(.;1) Γzcv,r(.;1) θzcv(.;1) ηzcv(.;1) Δzcv(.;1) Γzcv,Z(.;1) pde test
    0.01 59.3183 0.1474 0.0652 1.5974 0.0011 0.0524 0.0295 0.3782 1.33E17
    0.02 57.1534 0.2207 0.0814 1.6427 0.0012 0.0724 0.0383 0.4308 2.81E17
    0.03 55.0675 0.3103 0.0979 1.6404 0.0012 0.0946 0.0478 0.4781 3.42E18
    0.04 53.0577 0.4163 0.1141 1.5944 0.0012 0.1185 0.0578 0.5187 1.29E17
    0.05 51.1213 0.5382 0.1296 1.5102 0.0009 0.1437 0.0682 0.5515 2.04E17
    0.06 49.2555 0.6752 0.1442 1.3940 0.0006 0.1695 0.0787 0.5755 4.55E18
    0.07 47.4578 0.8261 0.1574 1.2523 0.0001 0.1954 0.0892 0.5902 1.68E17
    0.08 45.7258 0.9896 0.1691 1.0917 0.0004 0.2210 0.0995 0.5953 3.66E17
    0.09 44.0569 1.1639 0.1792 0.9186 0.0011 0.2458 0.1094 0.5907 1.91E17
    0.10 42.4490 1.3474 0.1875 0.7387 0.0019 0.2695 0.1188 0.5764 4.12E18
    0.11 40.8997 1.5383 0.1940 0.5571 0.0028 0.2917 0.1276 0.5528 2.91E17
    0.12 39.4070 1.7347 0.1986 0.3783 0.0037 0.3122 0.1356 0.5203 5.07E17
    0.13 37.9688 1.9350 0.2016 0.2060 0.0047 0.3308 0.1428 0.4794 1.65E17
    0.14 36.5830 2.1373 0.2028 0.0429 0.0058 0.3474 0.1491 0.4308 4.34E17
    0.15 35.2479 2.3400 0.2025 0.1087 0.0068 0.3620 0.1545 0.3750 9.97E18
    This table values 4-year call option prices (if α=1) and put option prices (if α=1), as well as their corresponding rho, interest rate gamma, theta, eta, delta, and gamma sensitivity measures, written on a 10-year zero-coupon bond with face value equal to $1.0 for different levels of the interest rate (r) and strike price K = $0.6, and assuming a CIR interest rate dynamics. Parameter values borrowed from Chan et al. (1992, Table Ⅲ): κ=0.2339, θ=0.0808, σ=0.0854, and λ=0. The zero-coupon bond prices and the option prices are both expressed as percentages of the face value. The last column of the table tests the CIR pde 12σ2rΓzcv,r+[κθ(κ+λ)r]ρzcv+θzcvrvzc=0 (the value on the left-hand-side is displayed to check how close it is to zero). The required noncentral chi-square distribution function and the corresponding probability density function have been computed using, respectively, the Benton and Krishnamoorthy (2003) algorithm and the built-in function ncx2pdf available in Matlab.

     | Show Table
    DownLoad: CSV

    A simple way to check the analytical formulas of our Greeks is to replace the solutions of the price, rho, interest rate gamma, and theta into the CIR partial differential equation (hereafter, pde). This exercise allowed us to conclude that the pde is satisfied in all the tested cases. It is noteworthy to recall that while symbolic algebra programs such as Mathematica or Maple can be used to derive Greeks through elementary differentiation—see, for example, Shaw (1998) who derives Greeks for stock options under the geometric Brownian motion assumption via Mathematica—, these novel analytical solutions are important for several reasons. Firstly, as argued by Carr (2001), the derivation of Greeks through symbolic algebra programs cannot replace an intuitive understanding of the role, genesis, and relationships between Greek measures. Secondly, as highlighted in Larguinho et al. (2013), the computation time needed for computing analytic Greeks is much smaller, which is of pivotal importance when a trader needs to design hedging strategies in real time. For instance, it takes only about 0.29 seconds to perform all the computations shown in Table 1. Thirdly, the existence of Greeks in closed-form allows its coding in any desired computer language, e.g., Matlab, Python, Fortran, R, or C. Lastly, as it will be shown in Section 4, the delta sensitivity measure can be used to price (and hedge) American-style options on zero-coupon bonds under the CIR model via the SHP pricing methodology developed by Chung and Shih (2009) and Ruas et al. (2013) for stock options.

    Table 2 adopts the parameters configuration of Wei (1997, Table Ⅱ) to value 5-year call option prices (if α=1) and put option prices (if α=1), as well as their corresponding rho, interest rate gamma, theta, eta, delta, and gamma sensitivity measures, written on a 15-year 10% coupon bond—with annual payment of the ten coupons to be delivered after the expiry date of the option contract—with face value equal to $100 for different levels of the interest rate (r), strike price K = $100, κ=0.25, θ=0.085, σ=0.05, and λ=0. The required noncentral chi-square distribution function has been computed via the Benton and Krishnamoorthy (2003) algorithm and the corresponding probability density function has been computed using the built-in function ncx2pdf available in Matlab.

    Table 2.  Prices and Greeks of European-style options on coupon-paying bonds under the CIR model.
    Call options
    r P(r,t,s) vcb(.;1) ρcbv(.;1) Γcbv,r(.;1) θcbv(.;1) ηcbv(.;1) Δcbv(.;1) Γcbv,Z(.;1) pde test
    0.04 126.1318 9.1833 92.5420 586.0740 1.3791 72.8855 30.2879 26.3717 4.38E17
    0.06 118.6380 7.4484 81.0065 567.9602 0.9106 67.1640 28.5330 33.5644 5.12E17
    0.08 111.6294 5.9407 69.8268 549.3779 0.5076 60.8831 26.4689 42.1126 3.90E17
    0.10 105.0732 4.6525 59.0753 524.2961 0.1782 54.0721 24.0988 51.1757 6.07E18
    0.12 98.9389 3.5737 48.9233 489.0691 0.0726 46.9028 21.4769 59.6656 2.95E17
    0.14 93.1981 2.6902 39.5845 443.1400 0.2452 39.6490 18.6999 66.4988 3.08E17
    0.16 87.8244 1.9836 31.2550 388.6802 0.3464 32.6228 15.8884 70.8323 9.11E18
    0.18 82.7931 1.4323 24.0685 329.5524 0.3880 26.1120 13.1658 72.2162 4.34E19
    0.20 78.0814 1.0129 18.0749 270.1006 0.3846 20.3330 10.6390 70.6356 2.17E18
    0.22 73.6678 0.7016 13.2408 214.1606 0.3514 15.4096 8.3860 66.4531 9.97E18
    0.24 69.5326 0.4762 9.4665 164.4774 0.3019 11.3740 6.4511 60.2875 6.94E18
    0.26 65.6572 0.3168 6.6099 122.5353 0.2466 8.1836 4.8466 52.8721 8.57E18
    0.28 62.0243 0.2067 4.5109 88.6897 0.1931 5.7450 3.5587 44.9264 8.02E18
    0.30 58.6179 0.1324 3.0114 62.4610 0.1456 3.9390 2.5561 37.0647 1.08E17
    Put options
    r P(r,t,s) vcb(.;1) ρcbv(.;1) Γcbv,r(.;1) θcbv(.;1) ηcbv(.;1) Δcbv(.;1) Γcbv,Z(.;1) pde test
    0.04 126.1318 0.0382 1.7847 63.3286 0.0217 1.5388 0.5841 7.4858 1.29E17
    0.06 118.6380 0.0885 3.3390 91.9180 0.0225 3.1537 1.1761 12.9249 1.77E17
    0.08 111.6294 0.1754 5.4324 116.1725 0.0044 5.5546 2.0592 19.5579 2.66E18
    0.10 105.0732 0.3084 7.9183 130.3677 0.0442 8.6996 3.2301 26.5295 1.80E17
    0.12 98.9389 0.4932 10.5569 131.1408 0.1319 12.4052 4.6344 32.7354 5.42E19
    0.14 93.1981 0.7299 13.0718 118.1960 0.2612 16.3866 6.1752 37.0747 0.00E+00
    0.16 87.8244 1.0135 15.2090 93.9454 0.4285 20.3208 7.7314 38.6850 1.71E17
    0.18 82.7931 1.3345 16.7814 62.4740 0.6247 23.9103 9.1796 37.0957 1.56E17
    0.20 78.0814 1.6803 17.6903 28.3308 0.8376 26.9291 10.4126 32.2691 2.78E17
    0.22 73.6678 2.0375 17.9239 4.4597 1.0544 29.2446 11.3520 24.5430 2.60E18
    0.24 69.5326 2.3931 17.5407 32.9780 1.2639 30.8162 11.9534 14.5092 1.04E17
    0.26 65.6572 2.7357 16.6445 55.5794 1.4576 31.6787 12.2043 2.8713 2.60E18
    0.28 62.0243 3.0563 15.3605 71.7604 1.6297 31.9177 12.1181 9.6831 6.94E18
    0.30 58.6179 3.3484 13.8147 81.8640 1.7778 31.6455 11.7259 22.5749 5.20E18
    This table values 5-year call option prices (if α=1) and put option prices (if α=1), as well as their corresponding rho, interest rate gamma, theta, eta, delta, and gamma sensitivity measures, written on a 15-year 10% coupon bond—with the ten coupons being paid annually—with face value equal to $100 for different levels of the interest rate (r) and strike price K = $100, and assuming a CIR interest rate dynamics. Parameter values borrowed from Wei (1997, Table Ⅱ): κ=0.25, θ=0.085, σ=0.05, and λ=0. We note that the obtained coupon-paying bond prices and the option prices are both expressed as percentages of the face value. All the Greeks reported in this table are 100 times the corresponding partial derivative of the option price, in order to be consistent with the rho values shown in Wei (1997, Table Ⅱ). The last column of the table tests the CIR pde 12σ2rΓcbv,r+[κθ(κ+λ)r]ρcbv+θcbvrvcb=0 (the value on the left-hand-side is displayed to check how close it is to zero). The required noncentral chi-square distribution function has been computed via the Benton and Krishnamoorthy (2003) algorithm. The corresponding probability density function has been computed using the built-in function ncx2pdf available in Matlab.

     | Show Table
    DownLoad: CSV

    Wei (1997, Table Ⅱ) reports only prices and rho values for calls. Column 3 of Table 2 reveals that our call option prices are similar to the ones presented in the third and forth columns of Wei (1997, Table Ⅱ, Panel A).13 Our interest rate deltas shown in column 4 of Table 2—obtained via equation (22) with α=1—are also similar to the ones presented in the third and forth columns of Wei (1997, Table Ⅱ, Panel B), which have been computed through a numerical integration scheme, as mentioned in Wei (1997, Footnote 9). For r24%, however, it seems that there are some text typos in Wei (1997, Table Ⅱ), because the corresponding absolute values are approximately equal.14 We recall that Cox et al. (1985) and Longstaff (1993) show that zero-coupon and coupon-paying bond call options are strictly decreasing functions of the interest rate. Thus, the first derivative of bond call options with respect to interest rates (i.e., interest rate deltas) illustrated in the third and fourth columns of Wei (1997, Table Ⅱ, Panel B) should always be negative, as shown in column 4 of our Table 2.

    13 If we use the Sankaran (1963) approximation for computing the noncentral chi-square distribution function we obtain exactly the same bond option prices reported in his third column.

    14 For example, when using his accurate method, Wei (1997, Table Ⅱ, Panel A) reports positive values for rho of 9.5254, 6.6349, 4.5098, and 2.9933 for r equal to 0.24, 0.26, 0.28, and 0.30, respectively, whereas we obtain the correct negative sign with values equal to −9.4665, −6.6099, −4.5109, and −3.0114, respectively. A similar pattern of wrong positive rho values is observed under his proposed approximate method for r0.24.

    To further check the analytical formulas of our Greeks, we have replaced the solutions for the price, rho, interest rate gamma, and theta into the pde of the problem, which is satisfied in all the tested cases. As a final testing exercise, we reproduce, in Table 3 (resp., Table 4), the computation of call and put deltas (resp., gammas) reported in Longstaff (1993, Tables 1 and 2). Again, our results are similar to the ones shown in Longstaff (1993, Tables 1 and 2), which, to the best of our knowledge, have been obtained via numerical methods or through standard symbolic derivation software since no analytical solutions of Greeks have been provided or mentioned in the paper. The use of our closed-form solutions provides accurate values for computing Greeks of coupon-paying bond options and with a very small computational burden. For instance, it takes only about 4.35 seconds to perform all the computations shown in Table 2.

    Table 3.  Deltas of European-style options on coupon-paying bonds under the CIR model.
    Panel A: 8% coupon bond in Longstaff (1993, Table 1)
    Call deltas Put deltas
    r K=960 K=980 K=1000 K=960 K=980 K = 1,000
    0.01 0.0456 0.0269 0.0120 0.0015 0.0004 0.0046
    0.02 0.0454 0.0267 0.0119 0.0015 0.0004 0.0047
    0.03 0.0452 0.0265 0.0118 0.0015 0.0003 0.0049
    0.04 0.0449 0.0263 0.0117 0.0015 0.0002 0.0050
    0.05 0.0447 0.0261 0.0116 0.0014 0.0001 0.0052
    0.06 0.0445 0.0259 0.0115 0.0014 0.0001 0.0053
    0.07 0.0442 0.0257 0.0113 0.0014 0.0000 0.0055
    0.08 0.0440 0.0256 0.0112 0.0014 0.0001 0.0056
    0.09 0.0438 0.0254 0.0111 0.0013 0.0002 0.0058
    0.10 0.0435 0.0252 0.0110 0.0013 0.0002 0.0060
    0.11 0.0433 0.0250 0.0109 0.0013 0.0003 0.0061
    0.12 0.0431 0.0248 0.0108 0.0013 0.0004 0.0063
    0.13 0.0428 0.0247 0.0107 0.0012 0.0004 0.0065
    0.14 0.0426 0.0245 0.0106 0.0012 0.0006 0.0066
    0.15 0.0424 0.0243 0.0105 0.0012 0.0007 0.0068
    Panel B: 14% coupon bond in Longstaff (1993, Table 2)
    Call deltas Put deltas
    r K=1,340 K=1,360 K=1,380 K=1,340 K=1,360 K=1,380
    0.01 0.0513 0.0373 0.0244 0.0014 0.0014 0.0001
    0.02 0.0511 0.0370 0.0242 0.0014 0.0013 0.0000
    0.03 0.0508 0.0368 0.0240 0.0014 0.0013 0.0001
    0.04 0.0506 0.0366 0.0239 0.0014 0.0013 0.0001
    0.05 0.0504 0.0364 0.0237 0.0014 0.0012 0.0002
    0.06 0.0501 0.0362 0.0235 0.0014 0.0012 0.0003
    0.07 0.0499 0.0360 0.0233 0.0014 0.0011 0.0004
    0.08 0.0496 0.0357 0.0232 0.0014 0.0011 0.0005
    0.09 0.0494 0.0355 0.0230 0.0014 0.0011 0.0006
    0.10 0.0492 0.0353 0.0228 0.0013 0.0010 0.0007
    0.11 0.0489 0.0351 0.0226 0.0013 0.0010 0.0007
    0.12 0.0487 0.0349 0.0225 0.0013 0.0009 0.0008
    0.13 0.0485 0.0347 0.0223 0.0013 0.0009 0.0009
    0.14 0.0482 0.0345 0.0221 0.0013 0.0008 0.0010
    0.15 0.0480 0.0343 0.0220 0.0013 0.0008 0.0011
    This table values 5-year call and put deltas on a 10-year 8% coupon bond (in Panel A) and 14% coupon bond (in Panel B)—with the ten coupons being paid annually—with par value 1,000 for different levels of the riskless interest rate (r) and strike price (K) assuming a CIR model. Parameter values borrowed from Longstaff (1993, Tables 1 and 2): κ=0.75, θ=0.08, σ2=0.014, and λ=0. The required noncentral chi-square distribution function has been computed via the Benton and Krishnamoorthy (2003) algorithm. The corresponding probability density function has been computed using the built-in function ncx2pdf available in Matlab.

     | Show Table
    DownLoad: CSV
    Table 4.  Gammas of European-style options on coupon-paying bonds under the CIR model.
    Panel A: 8% coupon bond in Longstaff (1993, Table 1)
    Call gammas Put gammas
    r K=960 K=980 K=1000 K=960 K=980 K = 1,000
    0.01 0.2482 0.1990 0.1195 0.0196 0.0743 0.1594
    0.02 0.2509 0.2007 0.1201 0.0206 0.0764 0.1626
    0.03 0.2535 0.2023 0.1207 0.0216 0.0785 0.1658
    0.04 0.2562 0.2039 0.1213 0.0226 0.0807 0.1691
    0.05 0.2589 0.2056 0.1219 0.0236 0.0829 0.1724
    0.06 0.2617 0.2072 0.1225 0.0247 0.0851 0.1758
    0.07 0.2644 0.2089 0.1231 0.0258 0.0874 0.1792
    0.08 0.2672 0.2106 0.1237 0.0270 0.0897 0.1827
    0.09 0.2700 0.2123 0.1244 0.0281 0.0921 0.1862
    0.10 0.2728 0.2140 0.1250 0.0293 0.0945 0.1898
    0.11 0.2757 0.2157 0.1256 0.0306 0.0970 0.1935
    0.12 0.2786 0.2174 0.1262 0.0318 0.0995 0.1972
    0.13 0.2815 0.2191 0.1268 0.0331 0.1021 0.2009
    0.14 0.2844 0.2208 0.1274 0.0345 0.1047 0.2047
    0.15 0.2873 0.2225 0.1280 0.0358 0.1073 0.2086
    Panel B: 14% coupon bond in Longstaff (1993, Table 2)
    Call gammas Put gammas
    r K=1,340 K=1,360 K=1,380 K=1,340 K=1,360 K=1,380
    0.01 0.1800 0.1635 0.1336 0.0066 0.0259 0.0586
    0.02 0.1820 0.1650 0.1347 0.0071 0.0269 0.0601
    0.03 0.1841 0.1667 0.1357 0.0076 0.0279 0.0617
    0.04 0.1861 0.1683 0.1368 0.0081 0.0289 0.0633
    0.05 0.1882 0.1699 0.1378 0.0087 0.0299 0.0649
    0.06 0.1903 0.1715 0.1389 0.0092 0.0310 0.0666
    0.07 0.1925 0.1732 0.1400 0.0098 0.0321 0.0683
    0.08 0.1946 0.1748 0.1410 0.0104 0.0332 0.0701
    0.09 0.1968 0.1765 0.1421 0.0110 0.0343 0.0718
    0.10 0.1990 0.1782 0.1432 0.0116 0.0355 0.0737
    0.11 0.2012 0.1799 0.1443 0.0122 0.0367 0.0755
    0.12 0.2034 0.1816 0.1454 0.0129 0.0379 0.0774
    0.13 0.2056 0.1833 0.1465 0.0135 0.0391 0.0793
    0.14 0.2079 0.1851 0.1476 0.0142 0.0404 0.0812
    0.15 0.2102 0.1868 0.1487 0.0149 0.0417 0.0832
    This table values 5-year call and put gammas on a 10-year 8% coupon bond (in Panel A) and 14% coupon bond (in Panel B)—with the ten coupons being paid annually—with par value 1,000 for different levels of the riskless interest rate (r) and strike price (K) assuming a CIR model. Parameter values borrowed from Longstaff (1993, Tables 1 and 2): κ=0.75, θ=0.08, σ2=0.014, and λ=0. The required noncentral chi-square distribution function has been computed via the Benton and Krishnamoorthy (2003) algorithm. The corresponding probability density function has been computed using the built-in function ncx2pdf available in Matlab.

     | Show Table
    DownLoad: CSV

    The goal now is to show how to implement the SHP approach for valuing American-style options on discount bonds under the CIR model. Even though the underlying asset is the bond, the independent variable is the stochastic interest rate r and there exists an unknown optimal exercise interest rate for which the exercise of the option becomes optimal. However, to provide a better understanding of the SHP method it is preferable to consider the unknown early exercise boundary, E, in terms of the underlying bond price Z. This requires the use of an alternative option pricing solution equivalent to equation (7), but expressed as a function of the underlying bond price Z.

    Let us first make a change of variable to express equation (7) as a function of the underlying bond price Z=Z(r,t,s).15 To accomplish this purpose, we note that

    15 We note that the presence of interim coupons prevents the use of a similar approach to equation (16) and, therefore, it is not possible to apply the SHP pricing methodology in the case of options on coupon-paying bonds.

    r(Z,t,s)=1B(t,s)ln(A(t,s)Z). (37)

    Substituting expression (37) into equations (7), (11), and (12), allows us to rewrite vzc(r,t,T,s,K;α)=ˉvzc(Z,t,T,s,K;α) as a function ˉvzc of Z instead of r and, therefore, rewrite expression (7) as

    ˉvzc(Z,t,T,s,K;α)=αZQ[x1(t,T,s,K);a,ˉb1(Z,t,T,s);α]αKA(t,T)(ZA(t,s))B(t,T)B(t,s)Q[x2(t,T,s,K);a,ˉb2(Z,t,T,s);α], (38)

    with

    ˉb1(Z,t,T,s):=2ϕ2(t,T)1B(t,s)lnA(t,s)Zeγ(Tt)ϕ(t,T)+ψ+B(T,s), (39)

    and

    ˉb2(Z,t,T,s):=2ϕ2(t,T)1B(t,s)lnA(t,s)Zeγ(Tt)ϕ(t,T)+ψ. (40)

    Armed with the alternative option pricing solution (38), we can now proceed with the derivation of the corresponding hedge ratio, i.e., the delta with respect to the underlying bond price Z.

    Proposition 8. Consider the pricing solution of a zero-coupon bond option under the CIR model as defined in equation (38). Then, the delta (with respect to the underlying bond price Z) of a zero-coupon bond call option (if α=1) or put option (if α=1) is given by

    Δzcˉv(.):=ˉvzc(Z,t,T,s,K;α)Z=αQ[x1(.);a,ˉb1(.);α]+2p(x1(.);a+2,ˉb1(.))ϕ2(t,T)eγ(Tt)B(t,s)[ϕ(t,T)+ψ+B(T,s)]αKA(t,T)B(t,T)ZB(t,s)(ZA(t,s))B(t,T)B(t,s)Q[x2(.);a,ˉb2(.);α]2KA(t,T)ZB(t,s)(ZA(t,s))B(t,T)B(t,s)p(x2(.);a+2,ˉb2(.))ϕ2(t,T)eγ(Tt)ϕ(t,T)+ψ. (41)

    Proof. Please see Appendix H.

    Remark 6. As expected,

    ˉvzc(Z,t,T,s,K;α)=vzc(r,t,T,s,K;α), (42)

    and

    Δzcˉv(Z,t,T,s,K;α)=Δzcv(r,t,T,s,K;α), (43)

    due to the relation between r and Z given by equation (2) or, equivalently, by equation (37). Note that this relation depends on t and s, but not on T and K. Hence, if we fix the values of t and s, the relation between r and Z is bijective and, therefore, working with one or the other to determine the price and the delta of the zero-coupon bond option is simply a (non-linear) scale change issue.16 Note also that r and Z are both random variables having exactly the same information, i.e., the same σ-algebra Ft. Therefore, by fixing t and s, Z is only a function of r (and, vice-versa, r is only a function of Z) so that we are able to use the classic chain rule for univariate functions in Remark 5. These arguments explain why we obtain the same values for prices and deltas of zero-coupon bond options using different (but equivalent) formulas.

    16 For the SHP approach, however, it is convenient to choose Z as the variable characterizing the underlying asset of the option.

    Let us define by ˉVzc(Z,t,T,s,K;α) the time-t price of an American-style call (if α=1) or put (if α=1) on the asset price Z (i.e., the underlying zero-coupon bond), with strike K, and maturity at time T (t).17 Let us denote the first passage time of the underlying asset price to its time-dependent exercise boundary {Eu,tuT} by

    17 Notice that the valuation of American-style call options on discount bonds can be performed via expressions (7) and (38), since such contracts, given the absence of interim coupons, will never be exercised before maturity—see Cox et al. (1985, Footnote 12). Nevertheless, for completeness, we will describe the SHP methodology for the general case.

    τ:=inf{u>t:Zu=Eu}. (44)

    Note that the critical asset price Zτ implies the existence of a critical interest rate rτ.

    Following Chung and Shih (2009) and Ruas et al. (2013), we use two well-known conditions on the early exercise boundary of the American option to solve this pricing problem: the value-matching and smooth-pasting conditions. At the maturity date T, if the American-style option has not been exercised earlier, its terminal condition is exactly the same as the corresponding European-style option. Therefore, we start at the maturity date of the American-style zero-coupon option and proceed backwards until the valuation date tt0. More specifically, at time T, we start our static hedge portfolio with one unit of the European-style zero-coupon bond option (38) with strike K, and expiry date at time T. In addition, we divide the time to maturity of the option contract into n evenly-spaced time points such that Δt:=(Tt0)/n. At each time ti:=t0+iΔt (for i=n1,,1,0), the unknown early exercise boundary Ei is matched by adding wi units of a standard European-style option with strike equal to Ei, and maturity at time ti+1. For each time step, the unknowns Ei and wi are found by solving simultaneously the following two value-matching and smooth-pasting recurrence conditions:

    αEniαK=ˉvzc(Eni,tni,T,s,K;α)+ij=1wnjˉvzc(Eni,tni,tnj+1,s,Enj;α), (45)

    and

    α=Δzcˉv(Eni,tni,T,s,K;α)+ij=1wnjΔzcˉv(Eni,tni,tnj+1,s,Enj;α), (46)

    for i=1,2,...,n, and with ˉvzc(.) and Δzcˉv(.) being given by expressions (38) and (41), respectively.

    After solving for all the unknowns Ei and wi (for i=n1,,1,0), the time-t0 SHP price of the American-style zero-coupon bond option, under the CIR model, is finally given by

    ˉVzc(Z0,t0,T,s,K;α):=ˉvzc(Z0,t0,T,s,K;α)+nj=1wnjˉvzc(Z0,t0,tnj+1,s,Enj;α). (47)

    Table 5 adopts the constellation of parameters used in Thakoor et al. (2012, Table 7), that is we consider a 5-year American-style put option on a 10-year zero-coupon bond with face value $100 for different levels of the interest rate (r) and an exercise price of $60. The CIR parameters are: κ=0.50, θ=0.08, σ=0.10, and λ=0. The third and fourth columns of the table report the European-style put prices computed via equations (7) and (38), respectively. Similarly, the sixth and seventh columns of the table show the results of put deltas obtained through equations (33) and (41), respectively. Columns 5 and 8 compute the corresponding differences and validate (numerically) the analytical equivalence demonstrated in Remark 6.

    Table 5.  Prices and deltas of European-style put options and prices of American-style put options on zero-coupon bonds under the CIR model.
    European puts European put deltas American puts
    r Z(r,t,s) vzc(.;1) ˉvzc(.;1) Diff. Δzcv(.;1) Δzcˉv(.;1) Diff n=2 n=100
    0.01 52.0729 0.0149 0.0149 8.67E19 0.0014 0.0014 2.10E16 7.9271 7.9271
    0.02 51.0671 0.0163 0.0163 1.73E18 0.0015 0.0015 6.94E18 8.9329 8.9329
    0.03 50.0807 0.0178 0.0178 8.67E19 0.0016 0.0016 2.67E16 9.9193 9.9193
    0.04 49.1134 0.0194 0.0194 1.73E18 0.0017 0.0017 1.24E16 10.8866 10.8866
    0.05 48.1647 0.0211 0.0211 8.67E19 0.0018 0.0018 2.36E16 11.8353 11.8353
    0.06 47.2344 0.0228 0.0228 0.00E+00 0.0019 0.0019 8.52E17 12.7656 12.7656
    0.07 46.3220 0.0246 0.0246 1.73E18 0.0020 0.0020 5.12E17 13.6780 13.6780
    0.08 45.4273 0.0265 0.0265 5.20E18 0.0022 0.0022 1.40E16 14.5727 14.5727
    0.09 44.5499 0.0284 0.0284 1.73E18 0.0023 0.0023 2.60E17 15.4501 15.4501
    0.10 43.6893 0.0304 0.0304 0.00E+00 0.0024 0.0024 3.37E16 16.3107 16.3107
    0.11 42.8455 0.0325 0.0325 1.73E18 0.0026 0.0026 2.75E16 17.1545 17.1545
    0.12 42.0179 0.0347 0.0347 1.73E18 0.0027 0.0027 2.91E16 17.9821 17.9821
    0.13 41.2063 0.0369 0.0369 5.20E18 0.0028 0.0028 3.17E17 18.7937 18.7937
    0.14 40.4104 0.0392 0.0392 5.20E18 0.0030 0.0030 1.80E16 19.5896 19.5896
    0.15 39.6298 0.0416 0.0416 5.20E18 0.0031 0.0031 1.28E16 20.3702 20.3702
    This table adopts the constellation of parameters used in Thakoor et al. (2012, Table 7), that is we consider a 5-year put option contract on a 10-year zero-coupon bond with face value $100 for different levels of the interest rate (r) and an exercise price of $60. The CIR parameters are: κ=0.50, θ=0.08, σ=0.10, and λ=0. The required noncentral chi-square distribution function has been computed via the Benton and Krishnamoorthy (2003) algorithm. The corresponding probability density function has been computed using the built-in function ncx2pdf available in Matlab.

     | Show Table
    DownLoad: CSV

    Regarding the American-style puts, we observe that the use of only two time-steps (i.e., n=2 in the SHP scheme) allow us to obtain the same price that is determined when using n=100. This implies that the static hedge portfolio replicating the American-style put requires only two European-style put options—at least for this combination of parameters—to hedge the American-style put. Finally, we note that, when rt=0.08, we are able to reproduce exactly the same put price (14.5727) reported in Thakoor et al. (2012, Table 7) when using both the Crank-Nicolson and the Jain's high-order compact schemes. In summary, the SHP approach can be viewed as a viable alternative to accurately and efficiently price American-style zero-coupon bond options under the CIR model.

    To further test the robustness of the proposed SHP scheme, Table 6 reports some additional numerical results by changing the parameters considered in Table 5. We observe, again, that the use of only two time-steps (i.e., n=2 in the SHP scheme) allow us to obtain the same price that is determined when using n=100. As already discussed in Cox et al. (1985) and Longstaff (1993), most of the comparative statics are indeterminate since changes in the interest rate parameters have complex effects on the relative values of bonds and options with different maturities. We note, however, that the observation that bond put prices can be decreasing functions of volatility is consistent with the counter effects arguments explained in Longstaff (1993, Pages 37–38).

    Table 6.  American-style put options on zero-coupon bonds under the CIR model.
    American puts
    κ θ σ Tt st K Z(r,t,s) n=2 n=100
    0.4 0.08 0.10 5.00 10.00 60 49.0169 10.9831 10.9831
    0.5 0.08 0.10 5.00 10.00 60 48.1647 11.8353 11.8353
    0.6 0.08 0.10 5.00 10.00 60 47.5871 12.4129 12.4129
    0.5 0.06 0.10 5.00 10.00 60 56.4233 3.5767 3.5767
    0.5 0.07 0.10 5.00 10.00 60 52.1307 7.8693 7.8693
    0.5 0.09 0.10 5.00 10.00 60 44.5005 15.4995 15.4995
    0.5 0.08 0.15 5.00 10.00 60 48.7285 11.2715 11.2715
    0.5 0.08 0.20 5.00 10.00 60 49.4724 10.5276 10.5276
    0.5 0.08 0.25 5.00 10.00 60 50.3605 9.6395 9.6395
    0.5 0.08 0.10 4.75 9.75 60 49.1168 10.8832 10.8832
    0.5 0.08 0.10 4.50 9.50 60 50.0874 9.9126 9.9126
    0.5 0.08 0.10 4.25 9.25 60 51.0768 8.9232 8.9232
    0.5 0.08 0.10 5.00 10.00 70 48.1647 21.8353 21.8353
    0.5 0.08 0.10 5.00 10.00 80 48.1647 31.8353 31.8353
    0.5 0.08 0.10 5.00 10.00 90 48.1647 41.8353 41.8353
    This table adopts the base case parameters considered in Table 5 and with rt=0.05. The CIR parameters κ, θ, and σ, the time to maturity of the put option (Tt), the time to maturity of the underlying bond (st), and the strike price of the put (K) are changed as shown in columns 1-6, respectivelly. The required noncentral chi-square distribution function has been computed via the Benton and Krishnamoorthy (2003) algorithm. The corresponding probability density function has been computed using the built-in function ncx2pdf available in Matlab.

     | Show Table
    DownLoad: CSV

    As expected, there might be some configurations of parameters where more time-steps are required. For instance, in Figure 1 we use the SHP method with n=8 and the following base case parameters borrowed from Yang (2004): a 1-year American-style put option on a 5-year zero-coupon bond with face value $100 for different levels of the interest rate (r) and an exercise price of $70. The CIR parameters are: κ=0.40, θ=0.08, σ=0.20, and λ=0. Then, we perform some numerical experiments by changing σ, t, κ, and θ. The graphs displayed in Figure 1 reveal the patterns that are expected under the CIR interest rate model and that are similar to the ones reported in Yang (2004, Figures 1 and 2). Hence, these additional results reported in Table 6 and Figure 1 corroborate the previous conclusion that the SHP approach is suitable for valuing American-style zero-coupon bond options under the CIR model and that a small number of time-steps in the SHP scheme is generally sufficient to obtain accurate values.

    Figure 1.  This figure plots 1-year American-style put options on a 5-year zero-coupon bond with face value $100 for different levels of the interest rate (r) and an exercise price of $70 borrowed from Yang (2004), using the proposed SHP method with n=8. CIR base case parameters: κ=0.40, θ=0.08, σ=0.20, and λ=0.
    Figure 2.  Stochastic duration of a sinking-fund bond under the CIR framework, for the same set of parameters as in Bacinello et al. (1996).

    Bonds are said to have embedded sinking-fund provisions when the issuer is required to retire portions of the bond issue before maturity, according to a pre-specified amortization schedule. The delivery option associated to this clause allows the issuer to retire the portions of the issue either by (ⅰ) calling the bonds by lottery at a pre-determinate value, usually at par, or (ⅱ) buying back the bonds at the prevailing market value.

    Bacinello et al. (1996) provide an elegant framework for analyzing the delivery option embedded in the sinking-fund bond provision (with only one sinking-fund date before maturity) under the one-dimensional stochastic term structure interest rate models of Vasicek (1977) and Cox et al. (1985). Bacinello et al. (1996) were able to analyze the comparative statics properties of the sinking-fund bond in the Vasicek (1977) framework analytically, but they use a numerical approach for the Cox et al. (1985) model. Thus, the main purpose of this section is to extend the Bacinello et al. (1996) approach by analyzing, in closed-form, the comparative statics properties of a default-free sinking-fund bond in the CIR framework.

    Following Bacinello et al. (1996), a sinking-fund bond is characterized by a coupon rate ic and an amortization schedule {(tj,Cj)}, where Cj>0 is the principal that the issuer is required to retire at time tj. We also assume that j=1,2 and, without loss of generality, C1+C2=1, i.e., the sinking-fund bond is issued with a normalized principal, retired in two dates only. Letting t0 denote the time of issuance of the bond, its coupon payments, Ij, are then assumed to be given by I1=(1+ic)(t1t0)1, and I2=C2[(1+ic)(t2t1)1]. At time t1 the issuer has the (delivery) option to retire the fraction C1 of the principal either by calling it by lottery at par value, or by buying it back at the market value.

    Bacinello et al. (1996, Proposition 2.1) show that the time-t price of the sinking-fund bond, Bsf(r,t), can be expressed either in terms of the corresponding serial bond and a bond put option, or in terms of the corresponding coupon bond and a bond call option, that is

    Bsf(r,t)=Bs(r,t)C1(1+ic)(t2t1)vzc(r,t,t1,t2,(1+ic)(t2t1);1), (48)

    or

    Bsf(r,t)=Bcb(r,t)C1(1+ic)(t2t1)vzc(r,t,t1,t2,(1+ic)(t2t1);1), (49)

    where Bs(r,t) and Bcb(r,t) represent, respectively, the time-t price of the corresponding serial and coupon bonds as given by Bacinello et al. (1996, Expressions 2.2 and 2.3).

    Let us now assume that t2t1=t1t0=1. Following the same line of reasoning applied by Bacinello et al. (1996) for the Vasicek (1977) framework, we substitute the relations given by Bacinello et al. (1996, Expressions 2.1 and 2.3) and the bond option pricing formula (7), with α=1, in expression (49). We then obtain, for t<t1,

    Bsf(r,t)=Z(r,t,t1)[ic+C1Q[x2(.);a,b2(.);1]]+(1+ic)Z(r,t,t2)[1C1Q[x1(.);a,b1(.);1]], (50)

    with x1(.) and x2(.) defined as in equations (8) and (9), but with K=(1+ic)1 in expression (15). Thus, the sinking-fund bond is shown to depend explicitly on the fraction C1 of outstanding capital to be retired at t1, the coupon rate ic, the spot rate r prevailing on the market, and the CIR parameters κ, θ, σ, and λ. We are now able to extend the analytical results provided by Bacinello et al. (1996) under the Vasicek (1977) framework for the CIR model case.

    The sinking-fund bond under the CIR model is an increasing function of the coupon rate. To establish this fact, take the derivative of (50) with respect to ic, and observe that the relation

    Z(r,t,t2)p(x1(.);a,b1(.))(ϕ(t,t1)+ψ+B(t1,t2))=Z(r,t,t1)p(x2(.);a,b2(.))(ϕ(t,t1)+ψ)(1+ic)1 (51)

    holds as an identity, so that, after some algebraic manipulations, we have

    Bsf(.)ic=Z(r,t,t1)+Z(r,t,t2)(1C1Q[x1(.);a,b1(.);1])>0, (52)

    where the strict positivity follows from the fact that, by assumption, 0<C1<1. Considering now the premiums Bcb(.)Bsf(.) and Bs(.)Bsf(.) of the corresponding coupon and serial bonds over the sinking-fund bond, and using respectively expressions (49) and (48), coupled with t2t1=t1t0=1, we obtain

    (Bcb(.)Bsf(.))ic=(Kvzc(.;1)vzc(.;1)K1)C1vzc(.;1)=C1Z(r,t,t2)Q[x1(.);a,b1(.);1]>0, (53)

    and

    (Bs(.)Bsf(.))ic=(Kvzc(.;1)vzc(.;1)K1)C1vzc(.;1)=C1Z(r,t,t2)Q[x1(.);a,b1(.);1]<0, (54)

    so that the higher the coupon rate, the larger is the premium demanded by the corresponding coupon bond over the sinking-fund bond, and the smaller is the premium determined by the corresponding serial bond over the sinking-fund bond. Note that the sign of the above derivatives depends entirely on the elasticity of the option prices to the strike price, in particular on the fact that such elasticity is negative for the call and exceeds 1 for the put.

    We can also explicitly analyze the comparative statics properties of the sinking-fund bond with respect to the spot rate r (rho) and time t (theta). The first sensitivity measure is given by

    ρsfB:=Bsf(.)r=Z(r,t,t1)r(ic+C1Q[x2(.);a,b2(.);1])+Z(r,t,t2)r(1+ic)(1C1Q[x1(.);a,b1(.);1])+C1(1+ic)Z(r,t,t2)p(x1(.);a+2,b1(.))2ϕ2(t,t1)eγ(tt1)ϕ(t,t1)+ψ+B(t1,t2)C1Z(r,t,t1)p(x2(.);a+2,b2(.))2ϕ2(t,t1)eγ(tt1)ϕ(t,t1)+ψ, (55)

    with Z(r,t,ti)/r, for i=t1,t2, given by equation (A.1). The effect on the premiums Bcb(.)Bsf(.) and Bs(.)Bsf(.) of an infinitesimal change in the spot interest rate r can be stated as

    (Bcb(.)Bsf(.))r=C1(1+ic)vzc(r,t,t1,t2,(1+ic)1;1)r, (56)
    (Bs(.)Bsf(.))r=C1(1+ic)vzc(r,t,t1,t2,(1+ic)1;1)r, (57)

    where vzc(r,t,t1,t2,(1+ic)1;α)/r is given by expression (21), with K=(1+ic)1.

    The effect on Bsf(.) of an infinitesimal change in t can be obtained explicitly as

    θsfB:=Bsf(.)t=Z(r,t,t1)t(ic+C1Q[x2(.);a,b2(.);1])+Z(r,t,t2)t(1+ic)(1C1Q[x1(.);a,b1(.);1])C1(1+ic)Z(r,t,t2)(p(x1(.);a,b1(.))ξp(x1(.);a+2,b1(.))ϱ1)+C1Z(r,t,t1)(p(x2(.);a,b2(.))ξp(x2(.);a+2,b2(.))ϱ2), (58)

    where Z(r,t,ti)/t, with i=t1,t2, is given by equation (C.3). As for the influence of the parameter t on the premiums Bcb(.)Bsf(.) and Bs(.)Bsf(.), we have

    (Bcb(.)Bsf(.))t=C1(1+ic)vzc(r,t,t1,t2,(1+ic)1;1)t, (59)
    (Bs(.)Bsf(.))t=C1(1+ic)vzc(r,t,t1,t2,(1+ic)1;1)t, (60)

    where vzc(r,t,t1,t2,(1+ic)1;α)/t is given by (25), with K=(1+ic)1.

    Now we want to prove a result that compares the stochastic durations of the sinking-fund bond with those of the corresponding serial and coupon bonds in the CIR model. Following Cox et al. (1979), the relative basis risk of a zero-coupon bond (under the CIR model), with maturity τ:=st, is given by g(τ)=2(eγτ1)/((κ+λ+γ)(eγτ1)+2γ)=B(t,s)=B(τ), a function that is strictly increasing (B(τ)/τ>0) and continuous on all positive reals, with the inverse function given by g1(τ)=(1/γ)ln(12γτ/((κ+λ+γ)τ2)), and defined on the interval ]0,2/(κ+λ+τ)[. Moreover, the stochastic duration of any interest rate sensitive instrument with price f(r,t) is given by

    Df=g1(x), (61)

    where x=(f(r,t)/r)/f(r,t) is the basis risk of f. Next proposition explicitly relates the stochastic durations of the sinking-fund, corresponding coupon and corresponding serial bonds under the CIR framework, thus extending the analytical results provided by Bacinello et al. (1996, Proposition 4.1), but for the Vasicek (1977) model.

    Proposition 9. For any set of parameters, the stochastic durations Dsf(r,t), Dcb(r,t), and Ds(r,t) of the sinking-fund, corresponding coupon and corresponding serial bonds under the CIR model satisfy the relation

    Ds(r,t)<Dsf(r,t)<Dcb(r,t). (62)

    Proof. Please see Appendix Ⅰ.

    Using the same set of parameters as in Bacinello et al. (1996), Figure 2 highlights that the stochastic duration of the sinking-fund bond is between the stochastic duration of the corresponding serial and coupon bonds. While this issue has been shown already by Bacinello et al. (1996, Figure 13) through numerical differentiation, we have now established this property analytically via Proposition 9 using the aforementioned novel closed-form solutions for the CIR Greeks.

    In this paper, we derive closed-form expressions for determining sensitivity measures of pure discount and coupon-paying bond options under the CIR framework, which are shown to be accurate, easy to implement, and computationally very efficient. The proposed hedge ratio allow us to evaluate American-style options on zero-coupon bonds through the static hedging approach. Moreover, we offer closed-form tractable expressions to analyze the comparative statics properties of a sinking-fund bond under the same interest rate dynamics setting.

    We thank two anonymous referees for their valuable comments and suggestions. Manuela Larguinho and Carlos A. Braumann belong to the research center CIMA (Centro de Investigação em Matemática e Aplicações, Instituto de Investigação e Formação Avançada, Universidade de Évora), supported by FCT (Fundação para a Ciência e a Tecnologia, Portugal), project UID/04674/2020. José Carlos Dias belongs to the Business Research Unit (BRU-IUL) and acknowledges the support provided by FCT [grant number UIDB/00315/2020].

    All authors declare no conflicts of interest in this paper.

    Let us first note that:

    Z(r,t,j)r=B(t,j)Z(r,t,j),j{T,s}, (A.1)
    x1(t,T,s,K)r=x2(t,T,s,K)r=0, (A.2)
    b1(r,t,T,s)r=b1(r,t,T,s)r, (A.3)

    and

    b2(r,t,T)r=b2(r,t,T)r. (A.4)

    The rho for a zero-coupon bond option is given by

    ρzcv(.):=vzc(.)r=αZ(r,t,s)rQ[x1(.);a,b1(.);α]+αZ(r,t,s)Q[x1(.);a,b1(.);α]rαK[Z(r,t,T)rQ[x2(.);a,b2(.);α]+Z(r,t,T)Q[x2(.);a,b2(.);α]r]. (A.5)

    Using expressions (17), (18), (A.2), (A.3), and (A.4) we are able to obtain the following partial derivatives:

    Q[x1(.);a,b1(.);α]r=Q[x1(.);a,b1(.);α]b1(.)b1(.)r=αb1(.)rp(x1(.);a+2,b1(.)), (A.6)

    and

    Q[x2(.);a,b2(.);α]r=Q[x2(.);a,b2(.);α]b2(.)b2(.)r=αb2(.)rp(x2(.);a+2,b2(.)). (A.7)

    Finally, substituting expressions (A.1), (A.6), and (A.7) into (A.5) yields expression (21).

    We first recall that Γzcv,r(.)=ρzcv(.)/r. Hence, differentiating (21) w.r.t. r and using (20), (A.1), (A.2), (A.3), (A.4), (A.6), and (A.7), expression (23) is finally obtained after straightforward calculations.

    Let us first note that:

    A(t,T)t=κθσ2(κ+λ+γ)(eγ(Tt)1)(2γ(κ+λ+γ))(κ+λ+γ)(eγ(Tt)1)+2γA(t,T), (C.1)
    B(t,T)t=4γ2eγ(Tt)[(κ+λ+γ)(eγ(Tt)1)+2γ]2, (C.2)

    and

    Z(r,t,j)t=Z(r,t,j)[1A(t,j)A(t,j)trB(t,j)t]=Z(r,t,j)ζj, (C.3)

    with

    ζj=1A(t,j)A(t,j)trB(t,j)t, (C.4)

    for j{T,s}.

    Let us also consider the following auxiliary functions:

    ξ=x1(t,T,s,K)t=x2(t,T,s,K)t=4rγ2eγ(Tt)σ2(eγ(Tt)1)2, (C.5)
    ϱ1=b1(r,t,T,s)t=b1(r,t,T,s)γ(ϕ(t,T)+ψ+B(T,s))+(ψ+B(T,s))eγ(Tt)(eγ(Tt)1)(ϕ(t,T)+ψ+B(T,s)), (C.6)

    and

    ϱ2=b2(r,t,T)t=b2(r,t,T)γ(ϕ(t,T)+ψ)+ψeγ(Tt)(eγ(Tt)1)(ϕ(t,T)+ψ). (C.7)

    The theta for a zero-coupon bond option is given by

    θzcv(.):=vzc(.)t=αZ(r,t,s)tQ[x1(.);a,b1(.);α]+αZ(r,t,s)Q[x1(.);a,b1(.);α]tαK[Z(r,t,T)tQ[x2(.);a,b2(.);α]+Z(r,t,T)Q[x2(.);a,b2(.);α]t]. (C.8)

    Using expressions (17), (18), (C.5), (C.6), and (C.7) we are able to compute the following partial derivatives:

    Q[x1(.);a,b1(.);α]t=Q[x1(.);a,b1(.);α)x1(.)x1(.)t+Q[x1(.);a,b1(.);α]b1(.)b1(.)t=αp(x1(.);a,b1(.))ξαp(x1(.);a+2,b1(.))ϱ1, (C.9)

    and

    Q[x2(.);a,b2(.);α]t=Q[x2(.);a,b2(.);α]x2(.)x2(.)t+Q[x2(.);a,b2(.);α]b2(.)b2(.)t=αp(x2(.);a,b2(.))ξαp(x2(.);a+2,b2(.))ϱ2. (C.10)

    Finally, substituting expressions (C.3), (C.9), and (C.10) into (C.8) yields expression (25).

    Let us first note that:

    x1(t,T,s,K)K=2(ϕ(t,T)+ψ+B(T,s))B(T,s)K, (D.1)
    x2(t,T,s,K)K=2(ϕ(t,T)+ψ)B(T,s)K, (D.2)

    and

    b1(r,t,T,s)K=b2(r,t,T)K=0. (D.3)

    The eta for a zero-coupon bond option is given by

    ηzcv(.):=vzc(.)K=αZ(r,t,s)Q[x1(.);a,b1(.);α]KαZ(r,t,T)[Q[x2(.);a,b2(.);α]+KQ[x2(.);a,b2(.);α]K]. (D.4)

    Using expressions (17), (18), (D.1), (D.2), and (D.3) we are able to compute the following partial derivatives:

    Q[(x1(.);a,b1(.);α]K=Q[x1(.);a,b1(.);α]x1(.)x1(.)K=2αp(x1(.);a,b1(.))(ϕ(t,T)+ψ+B(T,s))B(T,s)K, (D.5)

    and

    Q[x2(.);a,b2(.);α]K=Q[x2(.);a,b2(.);α]x2(.)x2(.)K=2αp(x2(.);a,b2(.))(ϕ(t,T)+ψ)B(T,s)K. (D.6)

    Finally, substituting expressions (D.5) and (D.6) into (D.4) yields expression (31).

    Let us first apply the decomposition technique of Jamshidian(1989) to obtain:

    ηcbv(.):=vcb(.)K=Ni=1aivzc(r,t,T,si,Ki;α)K. (E.1)

    Let us also compute, for an arbitrary fixed i, the expression for vzc(r,t,T,si,Ki;α)K=ziK, where zi=vzc(r,t,T,si,Ki;α). Note that the change of variable from K to r (keeping the remaining variables r, t, and T unchanged) is obtained as the implicit solution of K=Nj=1ajZ(r,T,sj) and so, applying the classical chain rule, one obtains

    vzc(r,t,T,si,Ki;α)K=ziK=zirrK=zir1Kr=zir1Nj=1ajZ(r,T,sj)r=zir1Nj=1aj[B(T,sj)Z(r,T,sj)]. (E.2)

    Moreover, with a new change of variable from r to Ki=Z(r,T,si) (keeping the remaining variables r, t, and T unchanged) and the application of the chain rule, leads to

    zir=ziKiKir. (E.3)

    Since

    Kir=Z(r,T,si)r=B(T,si)Z(r,T,si), (E.4)

    we obtain from (E.3) and (E.4),

    zir=ziKi[B(T,si)Z(r,T,si)]=vzc(r,t,T,si,Ki;α)Ki[B(T,si)Z(r,T,si)]=ηzcv(r,t,T,si,Ki;α)[B(T,si)Z(r,T,si)]. (E.5)

    Therefore, using (E.2) and (E.5),

    vzc(r,t,T,si,Ki;α)K=ηzcv(r,t,T,si,Ki;α)[B(T,si)Z(r,T,si)]1Nj=1aj[B(T,sj)Z(r,T,sj)] (E.6)

    and so, using (E.1) and (E.6), we obtain expression (32).

    Let us first note that:

    Γzcv,Z(.):=Δzcv(.)Z(r,t,s)=Δzcv(.)rrZ(r,t,s). (F.1)

    Using Remark 5, we conclude that

    rZ(r,t,s)=1B(t,s)Z(r,t,s). (F.2)

    Moreover,

    Δzcv(.)r=1B(t,s)r[ρzcv(.)Z(r,t,s)]=1B(t,s)[ρzcv(.)rZ(r,t,s)ρzcv(.)Z(r,t,s)r[Z(r,t,s)]2]. (F.3)

    Noting that ρzcv(.)/r=Γzcv,r(.), then expression (F.3) can be rewritten as

    Δzcv(.)r=1B(t,s)[Γzcv,r(.)Z(r,t,s)+ρzcv(.)B(t,s)Z(r,t,s)[Z(r,t,s)]2]=Γzcv,r(.)+ρzcv(.)B(t,s)B(t,s)Z(r,t,s). (F.4)

    Substituting expressions (F.2) and (F.4) into expression (F.1), we obtain

    Γzcv,Z(.)=Γzcv,r(.)+ρzcv(.)B(t,s)[B(t,s)Z(r,t,s)]2. (F.5)

    Finally, using expression (33) in equation (F.5) yields equation (35).

    Let us first note that:

    Γcbv,Z(.):=Δcbv(.)P(r,t,s)=Δcbv(.)rrP(r,t,s). (G.1)

    Using Remark 5, we conclude that

    rP(r,t,s)=1Ni=1aiB(t,si)Z(r,t,si). (G.2)

    Moreover,

    Δcbv(.)r=r[ρcbv(.)Ni=1aiB(t,si)Z(r,t,si)]=[ρcbv(.)rNi=1aiB(t,si)Z(r,t,si)ρcbv(.)Ni=1aiB(t,si)Z(r,t,si)(B(t,si))[Ni=1aiB(t,si)Z(r,t,si)]2]. (G.3)

    Noting that ρcbv(.)/r=Γcbv,r(.), then expression (G.3) can be rewritten as

    Δcbv(.)r=[Γcbv,r(.)Ni=1aiB(t,si)Z(r,t,si)+ρcbv(.)Ni=1ai[B(t,si)]2Z(r,t,si)[Ni=1aiB(t,si)Z(r,t,si)]2]. (G.4)

    Substituting expressions (G.2) and (G.4) into expression (G.1), we obtain

    Γcbv,Z(.)=Γcbv,r(.)Ni=1aiB(t,si)Z(r,t,si)+ρcbv(.)Ni=1ai[B(t,si)]2Z(r,t,si)[Ni=1aiB(t,si)Z(r,t,si)]3. (G.5)

    Finally, using expression (34) in equation (G.5) yields equation (36).

    Let us first note that:

    x1(t,T,s,K)Z=x2(t,T,s,K)Z=0, (H.1)
    ˉb1(Z,t,T,s)Z=2ϕ2(t,T)eγ(Tt)ZB(t,s)[ϕ(t,T)+ψ+B(T,s)], (H.2)

    and

    ˉb2(Z,t,T,s)Z=2ϕ2(t,T)eγ(Tt)ZB(t,s)[ϕ(t,T)+ψ]. (H.3)

    The delta for a zero-coupon bond option is computed as

    Δzcˉv(.):=ˉvzc(.)Z=αQ[x1(.);a,ˉb1(.);α]+αZQ[x1(.);a,ˉb1(.);α]ZαKA(t,T)A(t,s)B(t,T)B(t,s)(ZA(t,s))B(t,T)B(t,s)1Q[x2(.);a,ˉb2(.);α]αKA(t,T)(ZA(t,s))B(t,T)B(t,s)Q[x2(.);a,ˉb2(.);α]Z. (H.4)

    Using expressions (17), (18), (H.1), (H.2), and (H.3) we are able to obtain the following partial derivatives:

    Q[x1(.);a,ˉb1(.);α]Z=Q[x1(.);a,ˉb1(.);α]ˉb1(.)ˉb1(.)Z=2αp(x1(.);a+2,ˉb1(.))ϕ2(t,T)eγ(Tt)ZB(t,s)[ϕ(t,T)+ψ+B(T,s)], (H.5)

    and

    Q[x2(.);a,ˉb2(.);α]Z=Q[x2(.);a,ˉb2(.);α]ˉb2(.)ˉb2(.)Z=2αp(x2(.);a+2,ˉb2(.))ϕ2(t,T)eγ(Tt)ZB(t,s)[ϕ(t,T)+ψ]. (H.6)

    Finally, substituting expressions (H.5) and (H.6) into (H.4) yields expression (41).

    To verify the first inequality, we use expressions (48) and (55), along with the fact that g1(x) is (positive and) increasing, to observe that this inequality becomes

    1γln(12γρsB(κ+λ+γ)ρsB+2Bs(.))<1γln(12γρsfB(κ+λ+γ)ρsfB+2Bsf(.)),

    which is equivalent to vzc(r,t,t1,t2,(1+ic)1;1)ρsfBBs(r,t)ρzcp<0. To check the second inequality, we use now expression (49) and then follow the same reasoning to obtain Bcb(r,t)ρzccvzc(r,t,t1,t2,(1+ic)1;1)ρsfB<0, which concludes the proof.



    [1] Abramowitz M, Stegun IA (1972) Handbook of Mathematical Functions, (Dover, New York).
    [2] Allegretto W, Lin Y, Yang H (2003) Numerical Pricing of American Put Options on Zero-Coupon Bonds. Appl Numer Math 46: 113–134. https://doi.org/10.1016/S0168-9274(03)00034-5 doi: 10.1016/S0168-9274(03)00034-5
    [3] Alvarez LHR (2001) On the Form and Risk-Sensitivity of Zero Coupon Bonds for a Class of Interest Rate Models. Insur Math Econ 28: 83–90. https://doi.org/10.1016/S0167-6687(00)00068-8 doi: 10.1016/S0167-6687(00)00068-8
    [4] Bacinello AR, Ortu F, Stucchi P (1996) Valuation of Sinking-Fund Bonds in the Vasicek and CIR Frameworks. Appl Math Financ 3: 269–294. https://doi.org/10.1080/13504869600000013 doi: 10.1080/13504869600000013
    [5] Benton D, Krishnamoorthy K (2003) Computing Discrete Mixtures of Continuous Distributions: Noncentral Chisquare, Noncentral t and the Distribution of the Square of the Sample Multiple Correlation Coefficient. Comput Stat Data Anal 43: 249–267. https://doi.org/10.1016/S0167-9473(02)00283-9 doi: 10.1016/S0167-9473(02)00283-9
    [6] Brigo D, Mercurio F (2001) A Deterministic-Shift Extension of Analytically-Tractable and Time-Homogeneous Short-Rate Models. Financ Stoch 5: 369–387.
    [7] Carr P (2001) Deriving Derivatives of Derivative Securities. J Comput Financ 4: 5–29. https://10.1109/CIFER.2000.844609 doi: 10.1109/CIFER.2000.844609
    [8] Chan KC, Karolyi GA, Longstaff FA, et al. (1992) An Empirical Comparison of Alternative Models of the Short-Term Interest Rate. J Financ 47: 1209–1227. https://doi.org/10.1111/j.1540-6261.1992.tb04011.x doi: 10.1111/j.1540-6261.1992.tb04011.x
    [9] Chesney M, Elliott RJ, Gibson R (1993) Analytical Solutions for the Pricing of American Bond and Yield Options. Math Financ 3: 277–294. https://doi.org/10.1111/j.1467-9965.1993.tb00045.x doi: 10.1111/j.1467-9965.1993.tb00045.x
    [10] Chung SL, Shackleton M (2002) The Binomial Black-Scholes Model and the Greeks. J Futures Mark 22: 143–153. https://doi.org/10.1002/fut.2211 doi: 10.1002/fut.2211
    [11] Chung SL, Shackleton M (2005) On the Errors and Comparison of Vega Estimation Methods. J Futures Mark 25: 21–38. https://doi.org/10.1002/fut.20127 doi: 10.1002/fut.20127
    [12] Chung SL, Shih PT (2009) Static Hedging and Pricing American Options. J Bank Financ 33: 2140–2149. https://doi.org/10.1016/j.jbankfin.2009.05.016 doi: 10.1016/j.jbankfin.2009.05.016
    [13] Chung SL, Shih PT, Tsai WC (2010) A Modified Static Hedging Method for Continuous Barrier Options. J Futures Mark 30: 1150–1166. https://doi.org/10.1002/fut.20451 doi: 10.1002/fut.20451
    [14] Chung SL, Shih PT, Tsai WC (2013) Static Hedging and Pricing American Knock-In Put Options. J Bank Financ 37: 191–205. https://doi.org/10.1016/j.jbankfin.2012.08.019 doi: 10.1016/j.jbankfin.2012.08.019
    [15] Chung SL, Hung W, Lee HH, et al. (2011) On the Rate of Convergence of Binomial Greeks. J Futures Mark 31: 562–597. https://doi.org/10.1002/fut.20484 doi: 10.1002/fut.20484
    [16] Cohen JD (1988) Noncentral Chi-Square: Some Observations on Recurrence. Ame Stat 42: 120–122. https://10.1080/00031305.1988.10475540 doi: 10.1080/00031305.1988.10475540
    [17] Cox JC, Ingersoll Jr JE, Ross SA (1979) Duration and the Measurement of Basis Risk. J Bus 52: 51–61. http://www.jstor.org/stable/2352663
    [18] Cox JC, Ingersoll Jr JE, Ross SA (1985) A Theory of the Term Structure of Interest Rates. Econometrica 53: 385–408. https://doi.org/10.1142/9789812701022_0005 doi: 10.1142/9789812701022_0005
    [19] Cruz A, Dias JC (2017) The Binomial CEV Model and the Greeks. J Futures Mark 37: 90–104. https://doi.org/10.1002/fut.21791 doi: 10.1002/fut.21791
    [20] Cruz A, Dias JC (2020) Valuing American-Style Options under the CEV Model: An Integral Representation Based Method. Rev Deri Res 23: 63–83. https://doi.org/10.1007/s11147-019-09157-w doi: 10.1007/s11147-019-09157-w
    [21] Deng G (2015) Pricing American Put Option on Zero-Coupon Bond in a Jump-Extended CIR Model. Commun Nonlinear Sci 22: 186–196. https://doi.org/10.1016/j.cnsns.2014.10.003 doi: 10.1016/j.cnsns.2014.10.003
    [22] Dias JC, Nunes JPV, Cruz A (2020) A Note on Options and Bubbles under the CEV Model: Implications for Pricing and Hedging. Rev Deriv Res 23: 249–272. https://doi.org/10.1007/s11147-019-09164-x doi: 10.1007/s11147-019-09164-x
    [23] Dias JC, Nunes JPV, Ruas JP (2015) Pricing and Static Hedging of European-Style Double Barrier Options under the Jump to Default Extended CEV Model. Quant Financ 15: 1995–2010. https://doi.org/10.1080/14697688.2014.971049 doi: 10.1080/14697688.2014.971049
    [24] Feller W (1951) Two Singular Diffusion Problems. Annal Math 54: 173–182. https://doi.org/10.2307/1969318 doi: 10.2307/1969318
    [25] Guo JH, Chang LF (2020) Repeated Richardson Extrapolation and Static Hedging of Barrier Options under the CEV Model. J Futures Mark 40: 974–988. https://doi.org/10.1002/fut.22100 doi: 10.1002/fut.22100
    [26] Hull J, White A (1990) Valuing Derivative Securities Using the Explicit Finite Difference Method. J Financ Quant Anal 25: 87–100. https://doi.org/10.2307/2330889 doi: 10.2307/2330889
    [27] Jamshidian F (1989) An Exact Bond Option Formula. J Financ 44: 205–209. https://doi.org/10.1111/j.1540-6261.1989.tb02413.x doi: 10.1111/j.1540-6261.1989.tb02413.x
    [28] Jamshidian F (1995) A Simple Class of Square-Root Interest-Rate Models. Appl Math Financ 2: 61–72. https://doi.org/10.1080/13504869500000004 doi: 10.1080/13504869500000004
    [29] Johnson NL, Kotz S, Balakrishnan N (1995) Continuous Univariate Distributions, Vol. 2, 2nd ed. (John Wiley & Sons, New York).
    [30] Larguinho M, Dias JC, Braumann CA (2013) On the Computation of Option Prices and Greeks under the CEV Model. Quant Financ 13: 907–917. https://doi.org/10.1080/14697688.2013.765958 doi: 10.1080/14697688.2013.765958
    [31] Longstaff FA (1993) The Valuation of Options on Coupon Bonds. J Bank Financ 17: 27–42. https://doi.org/10.1016/0378-4266(93)90078-R doi: 10.1016/0378-4266(93)90078-R
    [32] Longstaff FA, Schwartz ES (2001) Valuing American Options by Simulation: A Simple Least-Squares Approach. Rev Financ Stud 14: 113–147. https://doi.org/10.1093/rfs/14.1.113 doi: 10.1093/rfs/14.1.113
    [33] Maghsoodi Y (1996) Solution of the Extended CIR Term Structure and Bond Option Valuation. Math Financ 6: 89–109. https://doi.org/10.1111/j.1467-9965.1996.tb00113.x doi: 10.1111/j.1467-9965.1996.tb00113.x
    [34] McKean Jr HP (1965) Appendix: A Free Boundary Problem for the Heat Equation Arising from a Problem of Mathematical Economics. Ind Manage Rev 6: 32–39.
    [35] Najafi AR, Mehrdoust F, Shirinpour S (2018) Pricing American Put Option On Zero-Coupon Bond Under Fractional CIR Model With Transaction Cost. Commun Stat Simulation Comput 47: 864–870. https://doi.org/10.1080/03610918.2017.1295153 doi: 10.1080/03610918.2017.1295153
    [36] Nawalkha SK, Beliaeva NA (2007) Efficient Trees for CIR and CEV Short Rate Models. J Altern Invest 10: 71–90. https://doi.org/10.3905/jai.2007.688995 doi: 10.3905/jai.2007.688995
    [37] Nelson DB, Ramaswamy K (1990) Simple Binomial Processes as Diffusion Approximations in Financial Models. Rev Financ Stud 3: 393–430. https://doi.org/10.1093/rfs/3.3.393 doi: 10.1093/rfs/3.3.393
    [38] Nunes JPV, Ruas JP, Dias JC (2015) Pricing and Static Hedging of American-Style Knock-in Options on Defaultable Stocks. J Bank Financ 58: 343–360. https://doi.org/10.1016/j.jbankfin.2015.05.003 doi: 10.1016/j.jbankfin.2015.05.003
    [39] Nunes JPV, Ruas JP, Dias JC (2020) Early Exercise Boundaries for American-Style Knock-Out Options, Eur J Oper Res 285: 753–766. https://doi.org/10.1016/j.ejor.2020.02.006 doi: 10.1016/j.ejor.2020.02.006
    [40] Pelsser A, Vorst TC (1994) The Binomial Model and the Greeks. J Deriv 1: 45–49. https://doi.org/10.3905/jod.1994.407888 doi: 10.3905/jod.1994.407888
    [41] Peng Q, Henry S (2018) On the Distribution of Extended CIR Model. Stat Pro Lett 142: 23–29. https://doi.org/10.1016/j.spl.2018.06.011 doi: 10.1016/j.spl.2018.06.011
    [42] Ruas JP, Dias JC, Nunes JPV (2013) Pricing and Static Hedging of American Options under the Jump to Default Extended CEV Model. J Bank Financ 37: 4059–4072. https://doi.org/10.1016/j.jbankfin.2013.07.019 doi: 10.1016/j.jbankfin.2013.07.019
    [43] Sankaran M (1963) Approximations to the Non-Central Chi-Square Distribution. Biometrika 50: 199–204. https://doi.org/10.2307/2333761 doi: 10.2307/2333761
    [44] Shaw W (1998) Modelling Financial Derivatives with Mathematica (Cambridge University Press, Cambridge, UK).
    [45] ShuJi L, ShengHong L (2006) Pricing American Interest Rate Option on Zero-Coupon Bond Numerically. Appl Math Comput 175: 834–850. https://doi.org/10.1016/J.AMC.2005.08.008 doi: 10.1016/J.AMC.2005.08.008
    [46] Thakoor N, Tangman Y, Bhuruth M (2012) Numerical Pricing of Financial Derivatives Using Jain's High-Order Compact Scheme. Math Sci 6: 1–16. https://doi.org/10.1186/2251-7456-6-72 doi: 10.1186/2251-7456-6-72
    [47] Tian Y (1992) A Simplified Binomial Approach to the Pricing of Interest-Rate Contingent Claims. J Financ Eng 1: 14–37.
    [48] Tian Y (1994) A Reexamination of Lattice Procedures for Interest Rate Contingent Claims. Adv Futures Options Res 7: 87–111. https://ssrn.com/abstract=5877
    [49] Vasicek O (1977) An Equilibrium Characterization of the Term Structure. J Financ Econ 5: 177–188. https://doi.org/10.1016/0304-405X(77)90016-2 doi: 10.1016/0304-405X(77)90016-2
    [50] Wei JZ (1997) A Simple Approach to Bond Option Pricing. J Futures Mark 17: 131–160.
    [51] Yang H (2004) American Put Options on Zero-Coupon Bonds and a Parabolic Free Boundary Problem. Int J Numer Anal Model 1: 203–215.
    [52] Zhou HJ, Yiu KFC, Li LK (2011) Evaluating American Put Options on Zero-Coupon Bonds by a Penalty Method. J Comput Appl Math 235: 3921–3931. https://doi.org/10.1016/j.cam.2011.01.038 doi: 10.1016/j.cam.2011.01.038
  • This article has been cited by:

    1. David Anderson, Urban Ulrych, Accelerated American option pricing with deep neural networks, 2023, 7, 2573-0134, 207, 10.3934/QFE.2023011
    2. 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, 2025, 9, 2573-0134, 300, 10.3934/QFE.2025010
  • Reader Comments
  • © 2022 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(3407) PDF downloads(352) Cited by(2)

Figures and Tables

Figures(2)  /  Tables(6)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog