MathIsimple
Back to Financial Mathematics
Course 8: Continuous-Time Models

Black-Scholes Option Pricing Model

The cornerstone of modern derivatives pricing: continuous-time stochastic models, closed-form solutions, and dynamic hedging with Greeks

Learning Objectives

  • 1.Understand geometric Brownian motion as a continuous-time stock price model
  • 2.Apply Itô's lemma to derive the Black-Scholes PDE via delta hedging
  • 3.Master the Black-Scholes formulas for European calls and puts
  • 4.Calculate and interpret the Greeks: Delta, Gamma, Vega, Theta, and Rho
  • 5.Understand implied volatility and the volatility smile phenomenon
  • 6.Apply Black-Scholes for practical option pricing and risk management

1. Continuous-Time Framework and Geometric Brownian Motion

Definition 1.1: Brownian Motion (Wiener Process)

A Brownian motion WtW_t is a continuous-time stochastic process with:

  • W0=0W_0 = 0
  • Independent increments: WtWsW_t - W_s is independent of past for t>st > s
  • Normal increments: WtWsN(0,ts)W_t - W_s \sim N(0, t-s)
  • Continuous paths: WtW_t is continuous in tt

Properties: E[Wt]=0\mathbb{E}[W_t] = 0, Var(Wt)=t\text{Var}(W_t) = t, Cov(Ws,Wt)=min(s,t)\text{Cov}(W_s, W_t) = \min(s,t)

Remark 1.1: Intuition Behind Brownian Motion

Think of Brownian motion as the limit of a random walk as time steps become infinitesimally small:

Wti=1nϵiΔt,ϵiN(0,1) i.i.d.W_t \approx \sum_{i=1}^{n} \epsilon_i \sqrt{\Delta t}, \quad \epsilon_i \sim N(0,1) \text{ i.i.d.}

Key feature: Quadratic variation [W]t=t[W]_t = t (not zero!) unlike deterministic functions.

Definition 1.2: Geometric Brownian Motion (GBM)

Stock price StS_t follows a geometric Brownian motion if:

dSt=μStdt+σStdWtdS_t = \mu S_t dt + \sigma S_t dW_t

where:

  • μ\mu is the drift (expected return)
  • σ\sigma is the volatility (standard deviation of returns)
  • dWtdW_t is the increment of Brownian motion
Theorem 1.1: Solution to GBM

The geometric Brownian motion has the explicit solution:

St=S0exp((μσ22)t+σWt)S_t = S_0 \exp\left(\left(\mu - \frac{\sigma^2}{2}\right)t + \sigma W_t\right)

Equivalently, log returns are normally distributed:

ln(St/S0)N((μσ22)t,σ2t)\ln(S_t/S_0) \sim N\left(\left(\mu - \frac{\sigma^2}{2}\right)t, \sigma^2 t\right)
Proof:

Let Yt=lnStY_t = \ln S_t. By Itô's lemma (to be stated formally later):

dYt=d(lnSt)=1StdSt121St2d[S]tdY_t = d(\ln S_t) = \frac{1}{S_t}dS_t - \frac{1}{2}\frac{1}{S_t^2}d[S]_t

Since d[S]t=σ2St2dtd[S]_t = \sigma^2 S_t^2 dt (quadratic variation):

dYt=1St(μStdt+σStdWt)12σ2dtdY_t = \frac{1}{S_t}(\mu S_t dt + \sigma S_t dW_t) - \frac{1}{2}\sigma^2 dt
dYt=(μσ22)dt+σdWtdY_t = \left(\mu - \frac{\sigma^2}{2}\right)dt + \sigma dW_t

Integrating from 0 to tt:

YtY0=(μσ22)t+σWtY_t - Y_0 = \left(\mu - \frac{\sigma^2}{2}\right)t + \sigma W_t
lnSt=lnS0+(μσ22)t+σWt\ln S_t = \ln S_0 + \left(\mu - \frac{\sigma^2}{2}\right)t + \sigma W_t

Exponentiating both sides yields the result.

Example 1.1: Stock Price Distribution

Setup: S0=$100S_0 = \$100, μ=0.10\mu = 0.10 (10% annual drift), σ=0.20\sigma = 0.20 (20% volatility), t=1t = 1 year

Expected value:

E[S1]=S0eμt=100e0.10$110.52\mathbb{E}[S_1] = S_0 e^{\mu t} = 100 e^{0.10} \approx \$110.52

Log return distribution:

ln(S1/S0)N(0.100.04/2,0.04)=N(0.08,0.04)\ln(S_1/S_0) \sim N(0.10 - 0.04/2, 0.04) = N(0.08, 0.04)

Probability S1>$120S_1 > \$120:

P(S1>120)=P(ln(S1/100)>ln(1.2))=P(Z>0.18230.080.2)P(Z>0.511)0.305P(S_1 > 120) = P(\ln(S_1/100) > \ln(1.2)) = P\left(Z > \frac{0.1823 - 0.08}{0.2}\right) \approx P(Z > 0.511) \approx 0.305

About 30.5% chance the stock exceeds $120 in one year.

Theorem 1.2: Itô's Lemma

Let XtX_t follow the SDE dXt=μtdt+σtdWtdX_t = \mu_t dt + \sigma_t dW_t and let f(t,Xt)f(t, X_t) be twice differentiable. Then:

df=(ft+μtfx+12σt22fx2)dt+σtfxdWtdf = \left(\frac{\partial f}{\partial t} + \mu_t \frac{\partial f}{\partial x} + \frac{1}{2}\sigma_t^2 \frac{\partial^2 f}{\partial x^2}\right)dt + \sigma_t \frac{\partial f}{\partial x} dW_t

The 12σ22fx2\frac{1}{2}\sigma^2 \frac{\partial^2 f}{\partial x^2} term (Itô correction) arises from the non-zero quadratic variation of Brownian motion.

Remark 1.2: Why Itô's Lemma Differs from Chain Rule

In deterministic calculus: df=f(x)dxdf = f'(x)dx

In stochastic calculus: (dWt)2=dt(dW_t)^2 = dt in the mean-square sense, not zero!

Taylor expansion to second order: dfftdt+fxdX+12fxx(dX)2df \approx f_t dt + f_x dX + \frac{1}{2}f_{xx}(dX)^2

Since (dX)2=σ2(dW)2=σ2dt(dX)^2 = \sigma^2(dW)^2 = \sigma^2 dt, the second-order term survives.

2. Deriving the Black-Scholes PDE

Definition 2.1: Black-Scholes Setup

Consider a European option with payoff Φ(ST)\Phi(S_T) at maturity TT. The option price V(t,St)V(t, S_t) is a function of time and stock price.

Assumptions:

  • Stock follows GBM: dSt=μStdt+σStdWtdS_t = \mu S_t dt + \sigma S_t dW_t
  • Risk-free rate rr is constant
  • No transaction costs, taxes, or dividends
  • Continuous trading possible
  • No arbitrage opportunities
Theorem 2.1: Black-Scholes Partial Differential Equation

The option price V(t,S)V(t,S) satisfies:

Vt+rSVS+12σ2S22VS2=rV\frac{\partial V}{\partial t} + rS\frac{\partial V}{\partial S} + \frac{1}{2}\sigma^2 S^2 \frac{\partial^2 V}{\partial S^2} = rV

with terminal condition V(T,S)=Φ(S)V(T, S) = \Phi(S).

Proof:

Step 1: Apply Itô's lemma to V(t,St)V(t, S_t):

dV=(Vt+μSVS+12σ2S22VS2)dt+σSVSdWtdV = \left(\frac{\partial V}{\partial t} + \mu S \frac{\partial V}{\partial S} + \frac{1}{2}\sigma^2 S^2 \frac{\partial^2 V}{\partial S^2}\right)dt + \sigma S \frac{\partial V}{\partial S} dW_t

Step 2: Construct a delta-hedged portfolio:

Π=VΔS\Pi = V - \Delta S

where Δ=VS\Delta = \frac{\partial V}{\partial S} is the hedge ratio. The portfolio change is:

dΠ=dVΔdSd\Pi = dV - \Delta dS
dΠ=(Vt+μSΔ+12σ2S22VS2)dt+σSΔdWtΔ(μSdt+σSdWt)d\Pi = \left(\frac{\partial V}{\partial t} + \mu S \Delta + \frac{1}{2}\sigma^2 S^2 \frac{\partial^2 V}{\partial S^2}\right)dt + \sigma S \Delta dW_t - \Delta(\mu S dt + \sigma S dW_t)

The stochastic terms cancel:

dΠ=(Vt+12σ2S22VS2)dtd\Pi = \left(\frac{\partial V}{\partial t} + \frac{1}{2}\sigma^2 S^2 \frac{\partial^2 V}{\partial S^2}\right)dt

Step 3: No-arbitrage argument:

Since Π\Pi is riskless (no dWtdW_t term), it must earn the risk-free rate:

dΠ=rΠdt=r(VΔS)dtd\Pi = r\Pi dt = r(V - \Delta S)dt

Equating the two expressions for dΠd\Pi:

Vt+12σ2S22VS2=r(VSVS)\frac{\partial V}{\partial t} + \frac{1}{2}\sigma^2 S^2 \frac{\partial^2 V}{\partial S^2} = r\left(V - S\frac{\partial V}{\partial S}\right)

Rearranging yields the Black-Scholes PDE.

Remark 2.1: Key Insights from the Derivation
  • Risk-neutral: The drift μ\mu disappeared! Option price doesn't depend on expected return
  • Dynamic hedging: Continuous rebalancing of Δ\Delta eliminates risk
  • Volatility matters: σ\sigma appears in the PDE and affects option value
  • Arbitrage-free: No-arbitrage implies the PDE; otherwise, arbitrage exists
Example 2.1: Verifying a Solution

Consider the simple case: V(t,S)=SV(t,S) = S (the stock itself as a "derivative").

Compute partial derivatives:

Vt=0,VS=1,2VS2=0\frac{\partial V}{\partial t} = 0, \quad \frac{\partial V}{\partial S} = 1, \quad \frac{\partial^2 V}{\partial S^2} = 0

Substitute into Black-Scholes PDE:

0+rS1+12σ2S20=rS0 + rS \cdot 1 + \frac{1}{2}\sigma^2 S^2 \cdot 0 = rS \quad \checkmark

The stock price itself satisfies the PDE (as it should)!

Proposition 2.1: Risk-Neutral Valuation Formula

The solution to the Black-Scholes PDE can be written as:

V(t,S)=er(Tt)EQ[Φ(ST)St=S]V(t,S) = e^{-r(T-t)} \mathbb{E}^Q[\Phi(S_T) \mid S_t = S]

where EQ\mathbb{E}^Q is expectation under the risk-neutral measure where:

dSt=rStdt+σStdWtQdS_t = rS_t dt + \sigma S_t dW_t^Q

Under QQ, the stock drifts at the risk-free rate rr (not the actual drift μ\mu).

3. Black-Scholes Formula for European Options

Theorem 3.1: Black-Scholes Formula

For a European call option with strike KK and maturity TT:

C(S,t)=SΦ(d1)Ker(Tt)Φ(d2)C(S, t) = S\Phi(d_1) - Ke^{-r(T-t)}\Phi(d_2)

For a European put option:

P(S,t)=Ker(Tt)Φ(d2)SΦ(d1)P(S, t) = Ke^{-r(T-t)}\Phi(-d_2) - S\Phi(-d_1)

where

d1=ln(S/K)+(r+σ2/2)(Tt)σTtd_1 = \frac{\ln(S/K) + (r + \sigma^2/2)(T-t)}{\sigma\sqrt{T-t}}
d2=d1σTt=ln(S/K)+(rσ2/2)(Tt)σTtd_2 = d_1 - \sigma\sqrt{T-t} = \frac{\ln(S/K) + (r - \sigma^2/2)(T-t)}{\sigma\sqrt{T-t}}

and Φ\Phi is the cumulative distribution function of the standard normal distribution.

Proof:

(Sketch) Under the risk-neutral measure, STLognormalS_T \sim \text{Lognormal}:

ST=Ste(rσ2/2)(Tt)+σTtZ,ZN(0,1)S_T = S_t e^{(r - \sigma^2/2)(T-t) + \sigma\sqrt{T-t}Z}, \quad Z \sim N(0,1)

Call price:

C=er(Tt)EQ[max(STK,0)]C = e^{-r(T-t)} \mathbb{E}^Q[\max(S_T - K, 0)]
C=er(Tt)max(Se(rσ2/2)(Tt)+σTtzK,0)ϕ(z)dzC = e^{-r(T-t)} \int_{-\infty}^{\infty} \max(Se^{(r-\sigma^2/2)(T-t) + \sigma\sqrt{T-t}z} - K, 0) \phi(z)dz

The call is in-the-money when ST>KS_T > K, i.e., z>d2z > -d_2. Evaluating the integral (using properties of log-normal distribution):

C=SΦ(d1)Ker(Tt)Φ(d2)C = S\Phi(d_1) - Ke^{-r(T-t)}\Phi(d_2)

The put formula follows from put-call parity or similar integration.

Example 3.1: Calculating Black-Scholes Call Price

Setup: S=$100S = \$100, K=$105K = \$105, r=0.05r = 0.05, σ=0.20\sigma = 0.20, Tt=1T - t = 1 year

Step 1: Calculate d1d_1 and d2d_2:

d1=ln(100/105)+(0.05+0.04/2)(1)0.201=0.0488+0.070.20=0.1062d_1 = \frac{\ln(100/105) + (0.05 + 0.04/2)(1)}{0.20\sqrt{1}} = \frac{-0.0488 + 0.07}{0.20} = 0.1062
d2=0.10620.20=0.0938d_2 = 0.1062 - 0.20 = -0.0938

Step 2: Look up normal CDF values:

Φ(0.1062)0.5423,Φ(0.0938)0.4626\Phi(0.1062) \approx 0.5423, \quad \Phi(-0.0938) \approx 0.4626

Step 3: Calculate call price:

C=100×0.5423105×e0.05×0.4626C = 100 \times 0.5423 - 105 \times e^{-0.05} \times 0.4626
C=54.23105×0.9512×0.4626=54.2346.19$8.04C = 54.23 - 105 \times 0.9512 \times 0.4626 = 54.23 - 46.19 \approx \$8.04
Example 3.2: Put Price via Put-Call Parity

Using data from Example 3.1, calculate the put price:

P=CS+Ker(Tt)P = C - S + Ke^{-r(T-t)}
P=8.04100+105×e0.05=8.04100+99.88$7.92P = 8.04 - 100 + 105 \times e^{-0.05} = 8.04 - 100 + 99.88 \approx \$7.92

Verification using put formula:

P=Ker(Tt)Φ(d2)SΦ(d1)P = Ke^{-r(T-t)}\Phi(-d_2) - S\Phi(-d_1)
P=99.88×(10.4626)100×(10.5423)P = 99.88 \times (1 - 0.4626) - 100 \times (1 - 0.5423)
P=99.88×0.5374100×0.4577=53.6745.77$7.90P = 99.88 \times 0.5374 - 100 \times 0.4577 = 53.67 - 45.77 \approx \$7.90

Slight discrepancy due to rounding; both methods give approximately $7.90-$7.92.

Remark 3.1: Interpreting the Black-Scholes Formula

Call formula: C=SΦ(d1)Ker(Tt)Φ(d2)C = S\Phi(d_1) - Ke^{-r(T-t)}\Phi(d_2)

  • Φ(d2)\Phi(d_2) ≈ risk-neutral probability that option expires in-the-money
  • Ker(Tt)Φ(d2)Ke^{-r(T-t)}\Phi(d_2) = discounted expected payment if exercised
  • Φ(d1)\Phi(d_1) = delta (hedge ratio); proportion of stock in replicating portfolio
  • SΦ(d1)S\Phi(d_1) = value of stock position in replicating portfolio

The formula decomposes the call into: value of stock received if exercised minus present value of strike paid.

Proposition 3.1: Limiting Cases

Deep in-the-money (SKS \gg K):

CSKer(Tt)(intrinsic value)C \to S - Ke^{-r(T-t)} \quad (\text{intrinsic value})

Deep out-of-the-money (SKS \ll K):

C0C \to 0

At-the-money (S=KS = K):

d1(r+σ2/2)(Tt)σTt,CS[Φ(d1)Φ(d2)]d_1 \approx \frac{(r + \sigma^2/2)(T-t)}{\sigma\sqrt{T-t}}, \quad C \approx S\left[\Phi(d_1) - \Phi(d_2)\right]

As maturity approaches (Tt0T - t \to 0):

Cmax(SK,0)(payoff)C \to \max(S - K, 0) \quad (\text{payoff})

4. The Greeks: Option Sensitivities

Definition 4.1: The Greeks

The Greeks measure the sensitivity of option prices to various parameters. They are essential for risk management and hedging.

Theorem 4.1: Delta (Δ)

Delta measures the rate of change of option price with respect to stock price:

Δ=VS\Delta = \frac{\partial V}{\partial S}

For Black-Scholes:

Δcall=Φ(d1),Δput=Φ(d1)1=Φ(d1)\Delta_\text{call} = \Phi(d_1), \quad \Delta_\text{put} = \Phi(d_1) - 1 = -\Phi(-d_1)

Interpretation: For every $1 increase in stock price, the option price increases by approximately Δ\Delta dollars.

Example 4.1: Delta Hedging

From Example 3.1, we have d1=0.1062d_1 = 0.1062, so Δcall=Φ(0.1062)0.54\Delta_\text{call} = \Phi(0.1062) \approx 0.54

To hedge a portfolio of 1000 short calls, buy 1000×0.54=5401000 \times 0.54 = 540 shares.

If stock rises by $1: Call loss ≈ 1000×0.54=$5401000 \times 0.54 = \$540, Stock gain = 540×1=$540540 \times 1 = \$540 → Net ≈ $0

Note: Delta changes as stock moves (see Gamma), requiring periodic rebalancing.

Theorem 4.2: Gamma (Γ)

Gamma measures the rate of change of Delta with respect to stock price:

Γ=2VS2=ΔS\Gamma = \frac{\partial^2 V}{\partial S^2} = \frac{\partial \Delta}{\partial S}

For Black-Scholes (same for call and put):

Γ=ϕ(d1)SσTt\Gamma = \frac{\phi(d_1)}{S\sigma\sqrt{T-t}}

where ϕ(x)=12πex2/2\phi(x) = \frac{1}{\sqrt{2\pi}}e^{-x^2/2} is the standard normal PDF.

Interpretation: Gamma measures the curvature of option value. High Gamma means Delta changes rapidly, requiring frequent rehedging.

Remark 4.1: Gamma and Delta-Hedging Costs

Gamma is highest for at-the-money options near expiry. This means:

  • Delta changes rapidly → frequent rebalancing needed
  • Higher transaction costs for maintaining delta-neutral portfolio
  • Gamma risk is significant for option sellers
Theorem 4.3: Vega (ν)

Vega measures sensitivity to volatility:

ν=Vσ\nu = \frac{\partial V}{\partial \sigma}

For Black-Scholes (same for call and put):

ν=Sϕ(d1)Tt\nu = S\phi(d_1)\sqrt{T-t}

Interpretation: For a 1% (0.01) increase in volatility, option price increases by ν×0.01\nu \times 0.01.

Example 4.2: Calculating Greeks

Using Example 3.1 data: S=$100S = \$100, K=$105K = \$105, r=0.05r = 0.05, σ=0.20\sigma = 0.20, Tt=1T-t = 1, d1=0.1062d_1 = 0.1062

Delta: Δ=Φ(0.1062)0.542\Delta = \Phi(0.1062) \approx 0.542

Gamma: ϕ(0.1062)=12πe0.10622/20.3977\phi(0.1062) = \frac{1}{\sqrt{2\pi}}e^{-0.1062^2/2} \approx 0.3977

Γ=0.3977100×0.20×1=0.3977200.0199\Gamma = \frac{0.3977}{100 \times 0.20 \times 1} = \frac{0.3977}{20} \approx 0.0199

Vega:

ν=100×0.3977×1=39.77\nu = 100 \times 0.3977 \times \sqrt{1} = 39.77

If volatility increases from 20% to 21%, call price increases by approximately 39.77×0.01$0.4039.77 \times 0.01 \approx \$0.40.

Theorem 4.4: Theta (Θ)

Theta measures time decay:

Θ=Vt\Theta = \frac{\partial V}{\partial t}

For a call:

Θcall=Sϕ(d1)σ2TtrKer(Tt)Φ(d2)\Theta_\text{call} = -\frac{S\phi(d_1)\sigma}{2\sqrt{T-t}} - rKe^{-r(T-t)}\Phi(d_2)

Typically Θ<0\Theta < 0 for long options: option value decreases as time passes (time decay).

Convention: Theta is often quoted as the change per day, so divide by 365.

Theorem 4.5: Rho (ρ)

Rho measures sensitivity to interest rate:

ρ=Vr\rho = \frac{\partial V}{\partial r}

For Black-Scholes:

ρcall=K(Tt)er(Tt)Φ(d2)\rho_\text{call} = K(T-t)e^{-r(T-t)}\Phi(d_2)
ρput=K(Tt)er(Tt)Φ(d2)\rho_\text{put} = -K(T-t)e^{-r(T-t)}\Phi(-d_2)

Interpretation: Typically the least important Greek for short-dated options, more relevant for long-dated options.

Remark 4.2: Greek Relationships

The Greeks satisfy the Black-Scholes PDE relationship:

Θ+rSΔ+12σ2S2Γ=rV\Theta + rS\Delta + \frac{1}{2}\sigma^2 S^2 \Gamma = rV

This means for a delta-hedged portfolio (Δ=0\Delta = 0):

Θ+12σ2S2Γ=rV\Theta + \frac{1}{2}\sigma^2 S^2 \Gamma = rV

Implication: Gamma profits offset Theta decay for a hedged position.

Proposition 4.1: Summary of Greeks for Calls and Puts
GreekCallPutTypical Sign
Delta (Δ)Φ(d1)\Phi(d_1)Φ(d1)1\Phi(d_1) - 1Call: +, Put: -
Gamma (Γ)ϕ(d1)SσTt\frac{\phi(d_1)}{S\sigma\sqrt{T-t}}Always +
Vega (ν)Sϕ(d1)TtS\phi(d_1)\sqrt{T-t}Always +
Theta (Θ)(Complex formula)Usually - (long)
Rho (ρ)>0> 0<0< 0Call: +, Put: -

5. Implied Volatility and the Volatility Smile

Definition 5.1: Implied Volatility

The implied volatility σimp\sigma_{\text{imp}} is the volatility parameter that, when input into the Black-Scholes formula, yields the observed market price:

Cmarket=CBS(S,K,r,Tt,σimp)C_{\text{market}} = C_{\text{BS}}(S, K, r, T-t, \sigma_{\text{imp}})

Since the Black-Scholes formula is monotonically increasing in σ\sigma, implied volatility exists and is unique for each option.

Remark 5.1: Computing Implied Volatility

There's no closed-form solution for σimp\sigma_{\text{imp}}. Common methods:

  • Newton-Raphson: Use Vega = Cσ\frac{\partial C}{\partial \sigma} for fast convergence
  • Bisection: Simple but slower, guaranteed convergence
  • Approximations: Various analytical approximations available
Example 5.1: Implied Volatility Calculation

Market data: Call with S=$100S = \$100, K=$100K = \$100, r=0.05r = 0.05, Tt=0.25T-t = 0.25 years trades at Cmkt=$5.50C_{\text{mkt}} = \$5.50

Newton-Raphson iteration: Start with σ0=0.30\sigma_0 = 0.30 (30%)

Iteration 1: CBS(σ0)=$5.23C_{\text{BS}}(\sigma_0) = \$5.23, ν=19.77\nu = 19.77

σ1=σ0+CmktCBS(σ0)ν=0.30+5.505.2319.770.314\sigma_1 = \sigma_0 + \frac{C_{\text{mkt}} - C_{\text{BS}}(\sigma_0)}{\nu} = 0.30 + \frac{5.50 - 5.23}{19.77} \approx 0.314

Iteration 2: CBS(0.314)=$5.50C_{\text{BS}}(0.314) = \$5.50

Implied volatility ≈ 31.4%. Converges in 2-3 iterations typically.

Definition 5.2: Volatility Smile and Skew

The volatility smile is the pattern where implied volatility varies across strike prices:

  • Smile: Higher IV for both deep ITM and OTM options (equity index options)
  • Skew: IV decreases as strike increases (individual equity options)
  • Smirk: Asymmetric pattern common in equity markets post-1987 crash

Implication: Black-Scholes assumption of constant volatility is violated in practice.

Remark 5.2: Causes of Volatility Smile

Several explanations for the smile:

  • Fat tails: Real returns have heavier tails than log-normal distribution
  • Leverage effect: Stock drops → higher leverage → higher volatility
  • Crash fears: Market participants price in tail risk (downside protection premium)
  • Supply/demand: Hedging demand affects option prices
Example 5.2: Volatility Skew Pattern

Typical equity index option IVs:

  • Strike 90% of spot: IV ≈ 25%
  • Strike 100% of spot (ATM): IV ≈ 20%
  • Strike 110% of spot: IV ≈ 18%

Interpretation: Out-of-the-money puts (downside protection) trade at higher IV, reflecting crash fears and hedging demand.

Proposition 5.1: Practical Implications

For traders:

  • Use implied volatility surfaces (IV as function of strike and maturity)
  • Volatility trading strategies exploit IV changes independent of direction
  • Greeks calculated with market-implied volatility, not historical

For modelers:

  • Black-Scholes is a benchmark; smile suggests need for extensions
  • Stochastic volatility models (Heston) or local volatility models address smile
  • Jump-diffusion models capture fat tails

Key Takeaways

  • 1.Geometric Brownian motion models stock prices with constant drift and volatility
  • 2.Itô's lemma extends calculus to stochastic processes, accounting for quadratic variation
  • 3.Black-Scholes PDE derived via delta hedging eliminates risk; option price independent of expected return μ
  • 4.Black-Scholes formula provides closed-form European option prices using risk-neutral valuation
  • 5.Greeks (Delta, Gamma, Vega, Theta, Rho) measure sensitivities essential for hedging and risk management
  • 6.Implied volatility and the volatility smile reveal market expectations and model limitations

Practice Problems

Black-Scholes Model
15
Questions
0
Correct
0%
Accuracy
1
Which of the following is NOT an assumption of the Black-Scholes model?
Not attempted
2
In the Black-Scholes formula, what does Φ(d₁) represent?
Not attempted
3
What is the key insight of Itô's lemma compared to standard calculus?
Not attempted
4
In the Black-Scholes PDE derivation, why does the drift μ disappear?
Not attempted
5
For an at-the-money call option, Delta is approximately:
Not attempted
6
Which Greek measures the option's sensitivity to volatility?
Not attempted
7
Gamma is highest for:
Not attempted
8
What is implied volatility?
Not attempted
9
The volatility smile suggests that:
Not attempted
10
For a delta-neutral portfolio, which relationship holds?
Not attempted
11
Under the risk-neutral measure Q, the stock drift is:
Not attempted
12
Which statement about Theta is TRUE?
Not attempted
13
The solution to dS = μS dt + σS dW is:
Not attempted
14
Put-call parity for European options states:
Not attempted
15
Which Greek is the same for both calls and puts with the same strike and maturity?
Not attempted

Frequently Asked Questions

What are the key assumptions of the Black-Scholes model?

(1) Stock price follows geometric Brownian motion with constant volatility, (2) No transaction costs or taxes, (3) Risk-free rate is constant, (4) No dividends, (5) Markets are efficient (no arbitrage), (6) European options only.

Why is the Black-Scholes formula so important?

It provides a closed-form solution for option prices, revolutionized derivatives markets, and introduced the concept of dynamic hedging. Despite its limitations, it remains the industry standard and foundation for more advanced models.

What is the most important Greek for hedging?

Delta is typically most important as it measures price sensitivity and is used for delta-neutral hedging. However, Gamma (curvature) and Vega (volatility risk) are also crucial for comprehensive risk management.

What is implied volatility?

Implied volatility is the volatility parameter that makes the Black-Scholes price equal to the market price. It represents the market's expectation of future volatility and varies across strikes (volatility smile/skew).

Can Black-Scholes be used for American options?

No, the closed-form formula only applies to European options. For American options, numerical methods (binomial trees, finite differences, Monte Carlo) are required, though American calls on non-dividend stocks equal European calls.