Skip to main content

Options (I): General Description, Parity Relations, Basic Concepts, and Valuation Using the Binomial Model

  • Chapter
  • First Online:
Capital Market Finance

Part of the book series: Springer Texts in Business and Economics ((STBE))

  • 1318 Accesses

Abstract

Option theory, which was developed at the start of the 1970s by Black, Scholes, and Merton, constituted a major advance in economic and financial theory. The applications of this theory extend well beyond its use for options. Not only do numerous financial products have option components (convertible bonds, caps and floors, hybrid products, …, and even the bonds and shares issued by limited companies where there is a risk of bankruptcy) but many decisions have an aspect that can only be understood in terms of options (investments, analysis of credit risk, etc.). Option theory provides tools that not only allow to price optional components but also to manage portfolios of assets and liabilities that may include them. By greatly improving our understanding of financial mechanisms and risk management, option theory has significantly contributed to the increase in activity on financial markets.

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

Access this chapter

eBook
USD 16.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Hardcover Book
USD 119.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    The possibility of filing for bankruptcy gives equity an optional aspect; this point of view is productive for analyzing corporate finance and credit risk, as we will see in Chaps. 28, 29, and 30.

  2. 2.

    Most investments allow future choices depending on the situation that will prevail; thus they include optional components which contribute to their net present value and which are termed “real options.”

  3. 3.

    We also will not be treating here the options called exotic, which are studied in Chap. 14, some of which have a payoff that depends on the path of the underlying’s price between 0 and T and not just on the value on date T (path dependent options such as American, Bermudian, on an average, barrier and lookback options).

  4. 4.

    As one will see in Fig. 10.6, the graph in Fig. 10.4 is relevant to the American put only.

  5. 5.

    Intuitively, the conditional probability of exercising the option Proba(ST < K | St)→ 0 as St→ ∞, so the expectation of the payoff E(YpT)→0 (since E(YpT< K Proba(ST < K | St)).

  6. 6.

    To simplify the notation, since to be completely explicit one would have to write: FT’(t).

  7. 7.

    We will denote by C, Ca, P, Pa, and S the values of options and their underlying on an arbitrary date t, except in the proof in Sect. 10.1.4.2 where we will be more specific and use Ct, Cat, Pt, Pat, and St.

  8. 8.

    In fact, case d does not necessarily correspond to a price drop since the only condition required of d is d < (1 + r) (see later on). However, since we do observe ups and downs in stock prices, we assume d < 1.

  9. 9.

    If the hedge portfolio’s value V0 is less than C0, it is sufficient, on date 0, to build this portfolio and sell the option (cashing in C0V0) and, on date 1, to settle the option’s payoff with what results from the sale of the hedge portfolio; this amounts to an arbitrage. In the opposite case (V0 > C0), taking the reverse position (buying the call and selling the portfolio) also is an arbitrage.

  10. 10.

    A dynamic strategy or portfolio is called self-financing if the investor does not disburse or withdraw funds on the intermediate dates i = 1, …, N–1. More precisely, on each date i = 1, …, N–1, the change in the portfolio’s composition leads to neither injection nor withdrawal of funds, the purchase of some assets being financed by the sale of others. Because an option is self-financing (there are no intermediate payments), the portfolio strategy that replicates it necessarily also is.

Suggestions for Further Reading

Books

  • Cox, J., & Rubinstein, M. (1985). Options markets. Prentice Hall.

    Google Scholar 

  • Hull, J. (2018). Options, futures and other derivatives (10th ed.). Prentice Hall Pearson Education.

    Google Scholar 

  • Kolb, R. (2007). Futures, options and swaps (5th ed.). Blackwell.

    Google Scholar 

  • McMillan, L. G. (1992). Options as a strategic investment. New York Institute of Finance.

    Google Scholar 

Articles

  • Cox, J., Ross, S., & Rubinstein, M. (1979). Option pricing, a simplified approach. Journal of Financial Economics, 7, 229–264.

    Article  Google Scholar 

  • Jarrow, R., & Rudd, A. (1982). Approximate option valuation for arbitrary stochastic processes. Journal of Financial Economics, 10, 347–369.

    Article  Google Scholar 

  • Stoll, H. (1969). The relationship between put and call option prices. Journal of Finance, 31, 319–332.

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Patrice Poncet .

Appendices

Appendix 1

1.1 Calibration of the Binomial Model

We start by computing the first three moments of a Bernoulli variable X that can take on two values u and d with respective probabilities q and 1 – q.

$$ {\displaystyle \begin{array}{l}E(X)= qu+\left(1\hbox{--} q\right)d\\ {}\mathit{\operatorname{var}}(X)=E\left({X}^2\right)\hbox{--} {\left[E(X)\right]}^2\\ {}={qu}^2+\left(1\hbox{--} q\right){d}^2\hbox{--} {q}^2{u}^2\hbox{--} {\left(1\hbox{--} q\right)}^2{d}^2+2q\left(1\hbox{--} q\right) ud\\ {}={u}^2q\left(1\hbox{--} q\right)+{d}^2\left(1\hbox{--} q\right)q+2q\left(1\hbox{--} q\right) ud,\end{array}} $$
(10.33)

so

$$ \operatorname{var}(X)=q\left(1-q\right){\left(u-d\right)}^2E{\left[X-E(X)\right]}^3=q{\left[u- qu-\left(1-q\right)d\right]}^3+\left(1-q\right)\ {\left[d- qu-\left(1-q\right)d\right]}^3=q{\left(1-q\right)}^3{\left(u-d\right)}^3+\left(1-q\right){q}^3{\left(d-u\right)}^3={\left(u-d\right)}^3\left[q\left(1-q\right)\left({\left(1-q\right)}^2-{q}^2\right)\right], $$
(10.34)

and it follows that

$$ E{\left[X\hbox{--} E(X)\right]}^3={\left(u\hbox{--} d\right)}^3q\left(1\hbox{--} q\left)\right(1\hbox{--} 2q\right). $$
(10.35)

We justify Calibration 2 (Jarrow and Rudd) given in Eqs. (10.32).

– On the one hand, in the Black-Scholes model, the logarithm of the stock price moves according to a Brownian process with independent increments of variance = σ2 and mean = \( \rho -\frac{\sigma^2}{2} \) per unit of time (i.e., σ2 T/N and (\( \rho -\frac{\sigma^2}{2} \))T/N for a period of length T/N). Furthermore, the third moment of each increment of lnS vanishes since it is a normal random variable (and hence has a symmetric distribution). Letting ΔilnS = ln(Si + 1) – ln(Si) (for the increment in an interval of duration T/N), we can therefore find binomial model parameters such that ΔilnS has the mean \( \left(\rho -\frac{1}{2}{\sigma}^2\right)\frac{T}{N} \) and the variance \( {\sigma}^2\frac{T}{N} \) (and, possibly, a vanishing third moment).

– The discretized binomial model means that ΔilnS = ln(Si + 1) – ln(Si) = ln(Si + 1/Si) is a binomial variable X taking on the two values lnuN and lndN with respective probabilities qN and 1 – qN. The first two moments of ΔilnS are thus given by Eqs. (10.33) and (10.34) (when we replace u and d by lnuN and lndN, and q by qN). Equate the first two moments of the binomial process with those of the Black-Scholes process:

$$ {q}_N\ln {u}_N+\left(1\hbox{--} {q}_N\right)\ln {d}_N=\left(\rho -\frac{1}{2}{\sigma}^2\right)\frac{T}{N} $$
(10.36)
$$ {q}_N\left(1-{q}_N\right){\left(\ln {u}_N-\ln {d}_N\right)}^2={\sigma}^2\frac{T}{N}. $$
(10.37)

An infinity of triples qN, lnuN, lndN can satisfy these two equations.

If we impose the additional condition that the third moment is zero, the three variables sought will have to obey a third equation:

$$ {\left(\ln {u}_N\hbox{--} \ln {d}_N\right)}^3\ {q}_N\ \left(1\hbox{--} {q}_N\right)\ \left(1\hbox{--} 2{q}_N\right)=0. $$

This last equation requires the choice qN =\( \frac{1}{2} \).

From this, lnuN and lndN are fixed by (10.36) and (10.37) with qN = 0.5, which gives

$$ \ln {u}_N=\left(\rho -\frac{1}{2}{\sigma}^2\right)\frac{T}{N}+\sigma \sqrt{\frac{T}{N}}\kern1em \mathrm{and}\kern1em \ln {d}_N=\left(\rho -\frac{1}{2}{\sigma}^2\right)\frac{T}{N}-\sigma \sqrt{\frac{T}{N}}. $$

(Without necessarily solving these equations, the reader could just check they are satisfied for these two values). The triple qN, lnuN, lndN is indeed that of Calibration 2 (Eqs. (10.32)).

Let us note that ln(SN) – ln(S0) = \( \sum \limits_{i=1}^N \) [ln(Si) – ln(Si–1)] is a sum of independent random variables that, by the central limit theorem, converges in law as N goes to infinity to a normal random variable with variance σ2 T and mean \( \left(\rho -\frac{1}{2}{\sigma}^2\right)T \). These are precisely the properties characteristic of the continuous Black-Scholes model (see the next chapter).

To finish, let us remark that for the values of uN and dN in Calibration 1 (Eqs. (10.31)), the variance of ΔilnS only equals \( {\sigma}^2\frac{T}{N} \) for qN = 0.5, which is not the exact value one has from Eqs. (10.31). However, the reader can check that the desired variance is attained asymptotically (since \( \underset{N\to \infty }{\lim }{q}_N \) = 0.5).

*Appendix 2

Proof of Proposition 12

(convergence of the Cox-Ross-Rubinstein formula to the Black-Scholes formula)

We assume that the binomial model is calibrated using Eqs. (10.32) (Calibration 2 justified in Appendix 1), start from the results of Proposition 12, and show that \( {\mathrm{Proba}}^{q_N}\left({S}_N\ge K\right)\to N\left({d}_2\right) \) and \( \Pr {\mathrm{oba}}^{q{\hbox{'}}_N}\left({S}_N\ge K\right)\to N\left({d}_1\right) \) if N → ∞, with \( {d}_1=\frac{\mathrm{In}\left({S}_0/K\right)+\left(\rho +{\sigma}^2/2\right)T}{\sigma \sqrt{T}} \) and \( {d}_2=\frac{\mathrm{In}\left({S}_0/K\right)+\left(\rho -{\sigma}^2/2\right)T}{\sigma \sqrt{T}}. \)

First, consider the value of SN as a function of the discrete random variable j, the number of increases during the period [0, T]; since SN = S0 uNj dNN–j and uN and dN are given by Eq. (10.32), we have

$$ {S}_N={S}_0{e}^{\left(\rho -\frac{1}{2}{\sigma}^2\right)T}{e}^{2 j\sigma \sqrt{\frac{T}{N}}- N\sigma \sqrt{\frac{T}{N}}}. $$

We deduce from this that the event SN ≥ K translates for j as

$$ j\ge \frac{\ln \left(K/{S}_0\right)-\left(\rho -\frac{1}{2}{\sigma}^2\right)T}{\sigma \sqrt{T}}\frac{\sqrt{N}}{2}+\frac{N}{2}. $$
(10.38)

Now let us examine the properties of the random variable j which is a binomial variable since it is the sum of N independent and identically distributed Bernoulli random variables. Indeed, define Xi for i = 1, …, N by: Xi = 1 if there is an increase between i and i + 1, and Xi = 0 otherwise.

Then \( {}_{i=1}^N{X}_i \) counts the number of increases occurring during the N periods between dates 0 and T, and we have

$$ j={\sum \limits}_{i=1}^N{X}_i. $$

The Central Limit Theorem applied to \( \frac{1}{N}{\sum \limits}_{i=1}^N{X}_i \) gives:

$$ \frac{\frac{1}{N}{\sum \limits}_{i=1}^N{X}_i-\frac{1}{N}E\left({}_{i=1}^N{X}_i\right)}{\frac{1}{N}\sigma \left({\sum \limits}_{i=1}^N{X}_i\right)}=\frac{{\sum \limits}_{i=1}^N{X}_i- NE\left({X}_i\right)}{\sqrt{N}\sigma \left({X}_i\right)}\to N\left(0,1\right). $$

In addition, E(Xi) = qN and σ2(Xi) = qN (1 – qN). We then have

$$ \frac{j-{Nq}_N}{\sqrt{Nq_N\left(1-{q}_N\right)}}\to N\left(0,1\right). $$

We will make \( \frac{j-{Nq}_N}{\sqrt{Nq_N\left(1-{q}_N\right)}} \) appear in Eq. (10.38):

$$ \frac{j-{Nq}_N}{\sqrt{Nq_N\left(1-{q}_N\right)}}\ge \left(\frac{\ln \frac{K}{S_0}-\left(\rho -\frac{1}{2}{\sigma}^2\right)T}{2\sigma \sqrt{T}}\sqrt{N}+\frac{N}{2}-{Nq}_N\right)\;\frac{1}{\sqrt{Nq_N\left(1-{q}_N\right)}}. $$
(10.39)

Let us use the calibration qN = 1/2 (Calibration 2, Eqs. (10.32)). We then obtain E(Xi) = qN = 1/2 and σ2(Xi) = qN (1 – qN) = 1/4. From this we have

$$ \frac{{\sum \limits}_{i=1}^N{X}_i- NE\left({X}_i\right)}{\sqrt{N}\sigma \left({X}_i\right)}=\frac{j-N/2}{\sqrt{N}/2}\to N\left(0,1\right). $$

Equation (10.39) becomes

$$ \frac{j-N/2}{\sqrt{N}/2}\ge \left(\left(\frac{\ln \frac{K}{S_0}-\left(\rho -\frac{1}{2}{\sigma}^2\right)T}{\sigma \sqrt{T}}\frac{\sqrt{N}}{2}+\frac{N}{2}\right)-\frac{N}{2}\right)\;\frac{1}{\sqrt{N}/2}=\left(\frac{\ln \frac{K}{S_0}-\left(\rho -\frac{1}{2}{\sigma}^2\right)T}{\sigma \sqrt{T}}\right)\equiv -{d}_2. $$
(10.40)

Thus we see that

$$ \Pr {\mathrm{oba}}^{q_N}\left({S}_N\ge K\right)=\Pr {\mathrm{oba}}^{q_N}\left(\frac{j-N/2}{\sqrt{N}/2}\ge -{d}_2\right)=\Pr \mathrm{oba}\left(U\ge -{d}_2\right) $$

where U is a reduced centered normal law under the probability being considered. N(u) is the standard notation for the distribution function of such a law:

$$ N(u)= Prob\left(U\le u\right). $$

By the law’s symmetry we have

$$ Prob\left(U\le {d}_2\right)= Prob\ \left(U\ge \hbox{--} {d}_2\right)=N\left({d}_2\right). $$

That finishes the proof of convergence for the second term of the call \( {C}_0^{(N)} \). In fact, we have shown

$$ \Pr {\mathrm{oba}}^{q_N}\left({S}_N\ge K\right)\to N\left({d}_2\right). $$

To prove that \( \Pr {\mathrm{oba}}^{q_N^{\prime }}\left({S}_N\ge K\right)\to N\left({d}_1\right) \), it is sufficient to repeat the reasoning while assigning q’N to the event Xi = 1 with \( {q}_N^{\prime }=\frac{q_N{u}_N}{e^{\rho T/N}}. \) Using Eq. (10.32) and a Taylor’s series expansion we obtain

$$ {q}_N^{\prime }=\frac{1}{2}\left(1+\sigma \sqrt{\frac{T}{N}}+o\left(\sqrt{\frac{T}{N}}\right)\right);q{'}_N\left(1\hbox{--} q{'}_N\right)=\frac{1}{4}\left(1+o\left(\sqrt{\frac{T}{N}}\right)\right). $$

So we have again inequality Eq. (10.39) with q’N instead of qN, by rewriting the second member as:

$$ \left(\left(\frac{\ln \frac{K}{S_0}-\left(\rho -\frac{1}{2}{\sigma}^2\right)T}{\sigma \sqrt{T}}\frac{\sqrt{N}}{2}+\frac{N}{2}\right)-\frac{N}{2}\left(1+\sigma \sqrt{\frac{T}{N}}+o\left(\sqrt{\frac{T}{N}}\right)\right)\right)\;\frac{2}{\sqrt{N}\sqrt{\left(1+o\left(\sqrt{\frac{T}{N}}\right)\right)}}=\left(\left(-{d}_2\frac{\sqrt{N}}{2}\right)-\frac{N}{2}\left(\sigma \sqrt{\frac{T}{N}}+o\left(\sqrt{\frac{T}{N}}\right)\right)\right)\;\frac{2}{\sqrt{N}\sqrt{\left(1+o\left(\sqrt{\frac{T}{N}}\right)\right)}}\kern6.1em =-{d}_2-\sigma \sqrt{T}+\sqrt{N}o\left(\sqrt{\frac{T}{N}}\right) $$

This term goes to \( -{d}_2-\sigma \sqrt{T} \) = − d1 as N tends to infinity, which proves the result.

The proof of convergence for a put can be inferred from this by using the call-put parity, or by replacing the event SN ≥ K by the event SN ≤ K.

Rights and permissions

Reprints and permissions

Copyright information

© 2022 The Author(s), under exclusive license to Springer Nature Switzerland AG

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Poncet, P., Portait, R. (2022). Options (I): General Description, Parity Relations, Basic Concepts, and Valuation Using the Binomial Model. In: Capital Market Finance. Springer Texts in Business and Economics. Springer, Cham. https://doi.org/10.1007/978-3-030-84600-8_10

Download citation

Publish with us

Policies and ethics