This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Asymptotic expansion for the Hartman-Watson distribution

Dan Pirjol School of Business
Stevens Institute of Technology
Hoboken, NJ 07030
[email protected]
(Date: 20 February 2021)
Abstract.

The Hartman-Watson distribution with density fr(t)=1I0(r)θ(r,t)f_{r}(t)=\frac{1}{I_{0}(r)}\theta(r,t) with r>0r>0 is a probability distribution defined on t0t\geq 0 which appears in several problems of applied probability. The density of this distribution is expressed in terms of an integral θ(r,t)\theta(r,t) which is difficult to evaluate numerically for small t0t\to 0. Using saddle point methods, we obtain the first two terms of the t0t\to 0 expansion of θ(ρ/t,t)\theta(\rho/t,t) at fixed ρ>0\rho>0. An error bound is obtained by numerical estimates of the integrand, which is furthermore uniform in ρ\rho. As an application we obtain the leading asymptotics of the density of the time average of the geometric Brownian motion as t0t\to 0. This has the form (1t0te2(Bs+μs)𝑑sda)(2πt)12g(a,μ)e1tJ(a)daa\mathbb{P}(\frac{1}{t}\int_{0}^{t}e^{2(B_{s}+\mu s)}ds\in da)\sim(2\pi t)^{-\frac{1}{2}}g(a,\mu)e^{-\frac{1}{t}J(a)}\frac{da}{a}, with an exponent J(a)J(a) which reproduces the known result obtained previously using Large Deviations theory.

Key words and phrases:
Asymptotic expansions, saddle point method, linear functional of the geometric Brownian motion

1. Introduction

The Hartman-Watson distribution was introduced in the context of directional statistics [15] and was studied further in relation to the first hitting time of certain diffusion processes [17]. This distribution has received considerable attention due to its relation to the law of the time integral of the geometric Brownian motion, see Eq. (3) below [25].

The Hartman-Watson distribution is given in terms of the function θ(r,t)\theta(r,t) defined as

(1) θ(r,t)=r2π3teπ22t0eξ22tercoshξsinhξsinπξtdξ.\theta(r,t)=\frac{r}{\sqrt{2\pi^{3}t}}e^{\frac{\pi^{2}}{2t}}\int_{0}^{\infty}e^{-\frac{\xi^{2}}{2t}}e^{-r\cosh\xi}\sinh\xi\sin\frac{\pi\xi}{t}d\xi\,.

The normalized function fr(t)=θ(r,t)I0(r)f_{r}(t)=\frac{\theta(r,t)}{I_{0}(r)} defines the density of a random variable taking values on the positive real axis t0t\geq 0, called the Hartman-Watson law [15].

The function θ(r,t)\theta(r,t) is related to the distribution of the time-integral of a geometric Brownian motion. Define

(2) At(μ)=0te2(Bs+μs)𝑑sA_{t}^{(\mu)}=\int_{0}^{t}e^{2(B_{s}+\mu s)}ds

with BtB_{t} a standard Brownian motion and μ\mu\in\mathbb{R}. The time-integral of the geometric Brownian motion is relevant for the pricing of Asian options in the Black-Scholes model [10], and appears also in actuarial science [5] and in the study of diffusion processes in random media [7]. Other applications to mathematical finance are related to the distributional properties of stochastic volatility models with log-normally distributed volatility, such as the β=1\beta=1 log-normal SABR model [1, 6] and the Hull-White model [13, 14].

An explicit result for the joint distribution of (At(μ),Bt)(A_{t}^{(\mu)},B_{t}) was given by Yor [25]

(3) (At(μ)du,Bt+μtdx)=eμx12μ2texp(1+e2x2u)θ(ex/u,t)dudxu,\displaystyle\mathbb{P}(A_{t}^{(\mu)}\in du,B_{t}+\mu t\in dx)=e^{\mu x-\frac{1}{2}\mu^{2}t}\exp\left(-\frac{1+e^{2x}}{2u}\right)\theta(e^{x}/u,t)\frac{dudx}{u}\,,

where the function θ(r,t)\theta(r,t) is given by (1).

The Yor formula (3) yields also the density of At(μ)A_{t}^{(\mu)} by integration over BtB_{t}. The usefulness of this approach is limited by issues with the numerical evaluation of the integral in (1) for small tt, due to the rapidly oscillating factor sin(πξ/t)\sin(\pi\xi/t) [3, 5]. Alternative numerical approaches which avoid this issue were presented in [4, 16].

For this reason considerable effort has been devoted to applying analytical methods to simplify Yor’s formula. For particular cases μ=0,1\mu=0,1 simpler expressions as single integrals have been obtained for the density of At(μ)A_{t}^{(\mu)} by Comtet et al [7] and Dufresne [9]. See the review article by Matsumoto and Yor [19] for an overview of the literature.

In the absence of simple exact analytical results, it is important to have analytical expansions of θ(r,t)\theta(r,t) in the small-tt region. Such an expansion has been derived by Gerhold in [12], by a saddle point analysis of the Laplace inversion integral of the density fr(t)f_{r}(t).

In this paper we derive the t0t\to 0 asymptotics of the function θ(r,t)\theta(r,t) at fixed rt=ρrt=\rho using the saddle point method. This regime is important for the study of the small-tt asymptotics of the density of 1tAt(μ)\frac{1}{t}A_{t}^{(\mu)} following from the Yor formula (3). The resulting expansion turns out to give also a good numerical approximation of θ(r,t)\theta(r,t) for all t0t\geq 0. The expansion has the form

(4) θ(ρ/t,t)=12πte1t(F(ρ)π22)(G(ρ)+G1(ρ)t+O(t2)).\theta(\rho/t,t)=\frac{1}{2\pi t}e^{-\frac{1}{t}\left(F(\rho)-\frac{\pi^{2}}{2}\right)}\left(G(\rho)+G_{1}(\rho)t+O(t^{2})\right)\,.

The leading order term proportional to G(ρ)G(\rho) is given in Proposition 1, and the subleading correction G1(ρ)G_{1}(\rho) is given in the Appendix.

The paper is structured as follows. In Section 2 we present the leading order asymptotic expansion of the integral θ(r,t)\theta(r,t) at fixed ρ=rt\rho=rt. The main result is Proposition 1. An explicit error bound for the leading term in the asymptotic expansion of θ(ρ/t,t)\theta(\rho/t,t) is given, guided by numerical tests, which is furthermore uniformly bounded in ρ\rho. In Section 3 we compare this asymptotic result with existing results in the literature on the small tt expansion of the Hartman-Watson distribution [3, 12]. In Section 4 we apply the expansion to obtain the leading t0t\to 0 asymptotics of the density of the time averaged geometric Brownian motion (1t0te2(Bs+μs)𝑑sda)=f(a,t)daa\mathbb{P}(\frac{1}{t}\int_{0}^{t}e^{2(B_{s}+\mu s)}ds\in da)=f(a,t)\frac{da}{a}. The leading asymptotics of this density has the form f(a,t)(2πt)12g(a,μ)e1tJ(a)f(a,t)\sim(2\pi t)^{-\frac{1}{2}}g(a,\mu)e^{-\frac{1}{t}J(a)}. The exponential factor J(a)J(a) reproduces a known result obtained previously using Large Deviations theory [2, 21]. An Appendix gives an explicit result for the subleading correction.

2. Asymptotic expansion of θ(ρ/t,t)\theta(\rho/t,t) as t0t\to 0

We study here the asymptotics of θ(r,t)\theta(r,t) as t0t\to 0 at fixed rt=ρrt=\rho. This regime is different from that considered in [12], which studied the asymptotics of θ(r,t)\theta(r,t) as t0t\to 0 at fixed rr.

Proposition 1.

The t0t\to 0asymptotics of the Hartman-Watson integral θ(ρ/t,t)\theta(\rho/t,t) is

(5) θ(ρ/t,t)=12πte1t(F(ρ)π22)(G(ρ)+G1(ρ)t+O(t2)),t0.\theta(\rho/t,t)=\frac{1}{2\pi t}e^{-\frac{1}{t}\left(F(\rho)-\frac{\pi^{2}}{2}\right)}\left(G(\rho)+G_{1}(\rho)t+O(t^{2})\right)\,,\quad t\to 0\,.

The function F(ρ)F(\rho) is given by

(9) F(ρ)={12x12ρcoshx1+π22,0<ρ<112y12+ρcosy1+πy1,ρ>1π221,ρ=1\displaystyle F(\rho)=\left\{\begin{array}[]{cc}\frac{1}{2}x_{1}^{2}-\rho\cosh x_{1}+\frac{\pi^{2}}{2}&\,,0<\rho<1\\ -\frac{1}{2}y_{1}^{2}+\rho\cos y_{1}+\pi y_{1}&\,,\rho>1\\ \frac{\pi^{2}}{2}-1&\,,\rho=1\\ \end{array}\right.

and the function G(ρ)G(\rho) is given by

(13) G(ρ)={ρsinhx1ρcoshx11,0<ρ<1ρsiny11+ρcosy1,ρ>13,ρ=1\displaystyle G(\rho)=\left\{\begin{array}[]{cc}\frac{\rho\sinh x_{1}}{\sqrt{\rho\cosh x_{1}-1}}&\,,0<\rho<1\\ \frac{\rho siny_{1}}{\sqrt{1+\rho\cos y_{1}}}&\,,\rho>1\\ \sqrt{3}&\,,\rho=1\\ \end{array}\right.

Here x1x_{1} is the solution of the equation

(14) ρsinhx1x1=1\rho\frac{\sinh x_{1}}{x_{1}}=1

and y1y_{1} is the solution of the equation

(15) y1+ρsiny1=π.y_{1}+\rho\sin y_{1}=\pi\,.

The subleading correction is G1(ρ)=12G(ρ)g~2(ρ)G_{1}(\rho)=\frac{1}{2}G(\rho)\tilde{g}_{2}(\rho) where g~2(ρ)\tilde{g}_{2}(\rho) is given in explicit form in the Appendix.

Plots of the functions F(ρ)F(\rho) and G(ρ)G(\rho) are shown in Figure 1. The properties of these functions are studied in more detail in Section 2.2.

Refer to caption
Refer to caption
Figure 1. Left: Plot of F(ρ)F(\rho) given in Eq. (9). The two branches in Eq. (9) are shown as the blue and red curves, respectively. The function F(ρ)F(\rho) has a minimum at ρ=π2\rho=\frac{\pi}{2}, with F(π2)=3π28F(\frac{\pi}{2})=\frac{3\pi^{2}}{8}. The dashed line shows the asymptotic line F(ρ)ρF(\rho)\sim\rho for ρ\rho\to\infty. Right: Plot of G(ρ)G(\rho) defined in Eq. (13).
Proof of Proposition 1.

The function θ(ρ/t,t)\theta(\rho/t,t) can be written as

(16) θ(ρ/t,t)=ρ/t2π3teπ22t0e1t[12ξ2+ρcoshξ]sinhξsinπξtdξ\displaystyle\theta(\rho/t,t)=\frac{\rho/t}{\sqrt{2\pi^{3}t}}e^{\frac{\pi^{2}}{2t}}\int_{0}^{\infty}e^{-\frac{1}{t}[\frac{1}{2}\xi^{2}+\rho\cosh\xi]}\sinh\xi\sin\frac{\pi\xi}{t}d\xi
=ρ/t2π3teπ22t12e1t[12ξ2+ρcoshξ]sinhξsinπξtdξ\displaystyle=\frac{\rho/t}{\sqrt{2\pi^{3}t}}e^{\frac{\pi^{2}}{2t}}\frac{1}{2}\int_{-\infty}^{\infty}e^{-\frac{1}{t}[\frac{1}{2}\xi^{2}+\rho\cosh\xi]}\sinh\xi\sin\frac{\pi\xi}{t}d\xi
=ρ2π3t3eπ22t14i(I+(ρ,t)I(ρ,t)),\displaystyle=\frac{\rho}{\sqrt{2\pi^{3}t^{3}}}e^{\frac{\pi^{2}}{2t}}\frac{1}{4i}(I_{+}(\rho,t)-I_{-}(\rho,t))\,,

with

(17) I±(ρ,t):=e1t[12ξ2+ρcoshξiπξ]sinhξdξ.I_{\pm}(\rho,t):=\int_{-\infty}^{\infty}e^{-\frac{1}{t}[\frac{1}{2}\xi^{2}+\rho\cosh\xi\mp i\pi\xi]}\sinh\xi d\xi\,.

These integrals have the form αβe1th(ξ)g(ξ)𝑑ξ\int_{\alpha}^{\beta}e^{-\frac{1}{t}h(\xi)}g(\xi)d\xi with h±(ξ)=12ξ2+ρcoshξiπξh_{\pm}(\xi)=\frac{1}{2}\xi^{2}+\rho\cosh\xi\mp i\pi\xi.

The asymptotics of I±(t)I_{\pm}(t) as t0t\to 0 can be obtained using the saddle-point method, see for example Sec. 4.6 in Erdélyi [11] and Sec. 4.7 of Olver [20].

We present in detail the asymptotic expansion for t0t\to 0 of the integral

(18) I+(ρ,t)=e1th(ξ)sinhξdξI_{+}(\rho,t)=\int_{-\infty}^{\infty}e^{-\frac{1}{t}h(\xi)}\sinh\xi d\xi

where we denote for simplicity h(ξ)=h+(ξ)=12ξ2+ρcoshξiπξh(\xi)=h_{+}(\xi)=\frac{1}{2}\xi^{2}+\rho\cosh\xi-i\pi\xi. The integral I(ρ,t)I_{-}(\rho,t) is treated analogously.

Refer to caption
Refer to caption
Refer to caption
Figure 2. Integration contours for I+(ρ,t)I_{+}(\rho,t) in the ξ\xi complex plane for the application of the asymptotic expansion. The contours for I(ρ,t)I_{-}(\rho,t) are obtained by changing the sign of yy (reflection in the real axis). The red dots show the saddle points. Left: contour for 0<ρ<10<\rho<1. The contour passes through the saddle points B(ξ=x1+iπ)B(\xi=-x_{1}+i\pi) and A(ξ=x1+iπ)A(\xi=x_{1}+i\pi). Middle: contour for ρ>1\rho>1. The contour passes through the saddle point S(ξ=iy1)S(\xi=iy_{1}). Right: the contour for ρ=1\rho=1 passes through the fourth order saddle point at S(ξ=iπ)S(\xi=i\pi).

i) 0<ρ<10<\rho<1. The saddle points are given by the solution of the equation h(ξ)=0h^{\prime}(\xi)=0. There are three saddle points at ξ={iπ,iπ±x1}\xi=\{i\pi,i\pi\pm x_{1}\} with x1x_{1} the solution of the equation x1ρsinhx1=0x_{1}-\rho\sinh x_{1}=0. The second derivative of hh at these points is h′′(iπ)=1ρ>0h^{\prime\prime}(i\pi)=1-\rho>0, and h′′(iπ±x1)=1ρcoshx1<0h^{\prime\prime}(i\pi\pm x_{1})=1-\rho\cosh x_{1}<0.

The contour of integration is deformed from the real axis to the contour shown in the left panel of Fig. 2, consisting of three arcs of curves of steepest descent passing through the three saddle points. Along these arcs we have Im[h(ξ)]=x(yπ)+ρsinhxsiny=0Im[h(\xi)]=x(y-\pi)+\rho\sinh x\sin y=0. Denoting ξ=x+iy\xi=x+iy, the path is given by

(21) y={π,|x|x1y0(x),|x|>x1\displaystyle y=\left\{\begin{array}[]{cc}\pi&\,,|x|\leq x_{1}\\ y_{0}(x)&\,,|x|>x_{1}\\ \end{array}\right.

where y0(x)y_{0}(x) is the positive solution of the equation ρsinhxsiny=(πy)x\rho\sinh x\sin y=(\pi-y)x.

The integral is a sum of three integrals along each piece of the path. The real part of the integrand is odd and the imaginary part is even under xxx\to-x. This follows from noting that we have Reh(x+iy)=Reh(x+iy)\mbox{Re}h(-x+iy)=\mbox{Re}h(x+iy) and sinh(x+iy)=sinhxcosy+isinycoshx\sinh(x+iy)=\sinh x\cos y+i\sin y\cosh x. This implies that i) the integral from B to A vanishes because the integrand is odd, and ii) the real parts of the integrals along (,B](-\infty,B] and [A,)[A,\infty) are equal and of opposite sign, and their imaginary parts are equal. This gives

(22) θ(ρ/t,t)=ρ2π3t3eπ22t12i(I(,B]+I[A,))=ρ2π3t3eπ22tImI[A,).\theta(\rho/t,t)=\frac{\rho}{\sqrt{2\pi^{3}t^{3}}}e^{\frac{\pi^{2}}{2t}}\frac{1}{2i}\left(I_{(-\infty,B]}+I_{[A,\infty)}\right)=\frac{\rho}{\sqrt{2\pi^{3}t^{3}}}e^{\frac{\pi^{2}}{2t}}\mbox{Im}I_{[A,\infty)}\,.

Thus it is sufficient to evaluate only ImI[A,)\mbox{Im}I_{[A,\infty)}. This integral is written as

(23) I[A,)(ρ,t)=Ae1th(ξ)sinhξdξ=e1th(A)0e1tτsinhξh(ξ)𝑑τ,I_{[A,\infty)}(\rho,t)=\int_{A}^{\infty}e^{-\frac{1}{t}h(\xi)}\sinh\xi d\xi=e^{-\frac{1}{t}h(A)}\int_{0}^{\infty}e^{-\frac{1}{t}\tau}\frac{\sinh\xi}{h^{\prime}(\xi)}d\tau\,,

where we defined τ=h(ξ)h(A)\tau=h(\xi)-h(A). This is expanded around ξ=A\xi=A as

(24) τ=h(ξ)h(A)=a2(ξA)2+a3(ξA)3+a4(ξA)4+O((ξA)5)\tau=h(\xi)-h(A)=a_{2}(\xi-A)^{2}+a_{3}(\xi-A)^{3}+a_{4}(\xi-A)^{4}+O((\xi-A)^{5})

with a2=12h′′(A),a3=16h′′′(A),a_{2}=\frac{1}{2}h^{\prime\prime}(A),a_{3}=\frac{1}{6}h^{\prime\prime\prime}(A),\cdots. Noting h′′(A)=1ρcoshx1<0h^{\prime\prime}(A)=1-\rho\cosh x_{1}<0, this series is inverted as

(25) ξA=i2τ|h′′(A)|+O(τ)\xi-A=-i\sqrt{\frac{2\tau}{|h^{\prime\prime}(A)|}}+O(\tau)

The second factor in the integrand of (23) is also expanded in ξA\xi-A as

(26) sinhξh(ξ)=sinhA+coshA(ξA)+O((ξA)2)h′′(A)(ξA)+O((ξA)2)=sinhAh′′(A)1ξA(1+O(ξA))\frac{\sinh\xi}{h^{\prime}(\xi)}=\frac{\sinh A+\cosh A(\xi-A)+O((\xi-A)^{2})}{h^{\prime\prime}(A)(\xi-A)+O((\xi-A)^{2})}=\frac{\sinh A}{h^{\prime\prime}(A)}\frac{1}{\xi-A}\left(1+O(\xi-A)\right)

Substituting here the expansion (25) gives

(27) sinhξh(ξ)=isinhx1|h′′(A)||h′′(A)|21τ+O(τ0)\frac{\sinh\xi}{h^{\prime}(\xi)}=i\frac{\sinh x_{1}}{|h^{\prime\prime}(A)|}\sqrt{\frac{|h^{\prime\prime}(A)|}{2}}\frac{1}{\sqrt{\tau}}+O(\tau^{0})

More generally

(28) g(τ)=sinhξh(ξ)=g0τ+g1+g2τ+O(τ).g(\tau)=\frac{\sinh\xi}{h^{\prime}(\xi)}=\frac{g_{0}}{\sqrt{\tau}}+g_{1}+g_{2}\sqrt{\tau}+O(\tau)\,.

The inequality h′′(A)<0h^{\prime\prime}(A)<0 implies that g0,g2,g_{0},g_{2},\cdots are imaginary, while g1,g3,g_{1},g_{3},\cdots are real.

By Watson’s lemma [20], the resulting expansion can be integrated term by term. The leading asymptotic contribution to I[A,)(ρ,t)I_{[A,\infty)}(\rho,t) is

(29) ImI[A,)(ρ,t)=e1th(A)Im0e1tτsinhξh(ξ)𝑑τ\displaystyle\mbox{Im}I_{[A,\infty)}(\rho,t)=e^{-\frac{1}{t}h(A)}\mbox{Im}\int_{0}^{\infty}e^{-\frac{1}{t}\tau}\frac{\sinh\xi}{h^{\prime}(\xi)}d\tau
=sinhx12|h′′(A)|e1th(A)(πt+O(t3/2)).\displaystyle\qquad=\frac{\sinh x_{1}}{\sqrt{2|h^{\prime\prime}(A)|}}e^{-\frac{1}{t}h(A)}\left(\sqrt{\pi t}+O(t^{3/2})\right)\,.

The leading order term integral was evaluated as 0e1tτdττ=πt\int_{0}^{\infty}e^{-\frac{1}{t}\tau}\frac{d\tau}{\sqrt{\tau}}=\sqrt{\pi t}. Substituting into (22) reproduces the quoted result by identifying h(A)=F(ρ)h(A)=F(\rho). Since g1,g3,g_{1},g_{3},\cdots are real, the O(τ0)O(\tau^{0}) term in (28) does not contribute to ImI[A,)ImI_{[A,\infty)}, and the leading correction comes from the O(τ1/2)O(\tau^{1/2}) term. This is given in explicit form in the Appendix.

ii) ρ>1\rho>1. There are several saddle points along the imaginary axis. We are interested in the saddle point at ξ=iy1\xi=iy_{1} with 0<y1π0<y_{1}\leq\pi the solution of (15). At this point the second derivative of hh is h′′(iy1)=1+ρcosy1>0h^{\prime\prime}(iy_{1})=1+\rho\cos y_{1}>0.

Deform the ξ:(,+)\xi:(-\infty,+\infty) integration contour into the curve in the middle panel of Fig.  2. This is a steepest descent curve Im(h(ξ))=0\mbox{Im}(h(\xi))=0, given by y0(x)y_{0}(x), the positive solution of the equation ρsinhxsiny=(πy)x\rho\sinh x\sin y=(\pi-y)x. The contour passes through the saddle point SS at ξ=iy1\xi=iy_{1}.

The integral I+(ρ,t)=S+S+I_{+}(\rho,t)=\int_{-\infty}^{S}+\int_{S}^{+\infty} is the sum of the two integrals on the two halves of the contour. As in the previous case, the sum is imaginary since h(ξ)h(\xi) is real along the contour, and Re[h(ξ)]Re[h(\xi)] is even in xx. Noting that sinh(x+iy)=sinhxcosy+isinycoshx\sinh(x+iy)=\sinh x\cos y+i\sin y\cosh x, the first term is odd in xx and cancels out, and only the second (imaginary) term gives a contribution. We have a result similar to (22)

(30) θ(ρ/t,t)=ρ2π3t3eπ22t12i(I(,S]+I[S,))=ρ2π3t3eπ22tImI[S,).\theta(\rho/t,t)=\frac{\rho}{\sqrt{2\pi^{3}t^{3}}}e^{\frac{\pi^{2}}{2t}}\frac{1}{2i}\left(I_{(-\infty,S]}+I_{[S,\infty)}\right)=\frac{\rho}{\sqrt{2\pi^{3}t^{3}}}e^{\frac{\pi^{2}}{2t}}\mbox{Im}I_{[S,\infty)}\,.

As before, it is sufficient to evaluate only the [S,)[S,\infty) integral, which is

(31) I[S,)(ρ,t)=Se1th(ξ)sinhξdξ=e1th(S)0e1tτsinhξh(ξ)𝑑τI_{[S,\infty)}(\rho,t)=\int_{S}^{\infty}e^{-\frac{1}{t}h(\xi)}\sinh\xi d\xi=e^{-\frac{1}{t}h(S)}\int_{0}^{\infty}e^{-\frac{1}{t}\tau}\frac{\sinh\xi}{h^{\prime}(\xi)}d\tau

where we introduced τ=h(ξ)h(S)0\tau=h(\xi)-h(S)\geq 0. This difference is expanded around ξ=S\xi=S as

(32) τ=h(ξ)h(S)=12h′′(S)(ξS)2+O((ξS)3)\tau=h(\xi)-h(S)=\frac{1}{2}h^{\prime\prime}(S)(\xi-S)^{2}+O((\xi-S)^{3})

which is inverted as (recall h′′(S)>0h^{\prime\prime}(S)>0)

(33) ξS=2τh′′(S)+O(τ)\xi-S=\sqrt{\frac{2\tau}{h^{\prime\prime}(S)}}+O(\tau)

The integrand is also expanded in ξS\xi-S as

(34) sinhξh(ξ)=sinhS+coshS(ξS)+O((ξS)2)h′′(S)(ξS)+O((ξS)2)=sinhSh′′(S)1ξS(1+O((ξS))).\frac{\sinh\xi}{h^{\prime}(\xi)}=\frac{\sinh S+\cosh S(\xi-S)+O((\xi-S)^{2})}{h^{\prime\prime}(S)(\xi-S)+O((\xi-S)^{2})}=\frac{\sinh S}{h^{\prime\prime}(S)}\frac{1}{\xi-S}\left(1+O((\xi-S))\right)\,.

Substituting here (33) this can be expanded in powers of τ\sqrt{\tau} as

(35) sinhξh(ξ)=g0τ+g1+g2τ+O(τ).\frac{\sinh\xi}{h^{\prime}(\xi)}=\frac{g_{0}}{\sqrt{\tau}}+g_{1}+g_{2}\sqrt{\tau}+O(\tau)\,.

As in the previous case, g0,g2,g_{0},g_{2},\cdots are imaginary, while g1,g3,g_{1},g_{3},\cdots are real, such that only the even order terms contribute.

By Watson’s lemma [20] we can exchange expansion and integration, and we get the leading asymptotic term

(36) Im0e1tτsinhξh(ξ)𝑑τ=siny12h′′(S)(πt+O(t3/2)).Im\int_{0}^{\infty}e^{-\frac{1}{t}\tau}\frac{\sinh\xi}{h^{\prime}(\xi)}d\tau=\frac{\sin y_{1}}{\sqrt{2h^{\prime\prime}(S)}}\left(\sqrt{\pi t}+O(t^{3/2})\right)\,.

The O(t3/2)O(t^{3/2}) subleading term is given in explicit form in the Appendix.

iii) ρ=1\rho=1. Some details of proof are different for ρ=1\rho=1 since the saddle point SS at ξ=iπ\xi=i\pi is of fourth order, so the Taylor expansion of τ\tau starts with the fourth order term in ξS\xi-S

(37) τ=h(ξ)h(S)=14!h′′′′(S)(ξS)4+,h′′′′(S)=1\tau=h(\xi)-h(S)=\frac{1}{4!}h^{\prime\prime\prime\prime}(S)(\xi-S)^{4}+\cdots\,,\quad h^{\prime\prime\prime\prime}(S)=-1

This expansion in z:=ξSz:=\xi-S can be put in closed form as

(38) τ=14!z416!z6+O(z8)=coshz+1+12z2.\tau=-\frac{1}{4!}z^{4}-\frac{1}{6!}z^{6}+O(z^{8})=-\cosh z+1+\frac{1}{2}z^{2}\,.

The inversion of this expansion gives

(39) z=ξS\displaystyle z=\xi-S =\displaystyle= (24τ)1/41120(24τ)3/4+1167200(24τ)5/4+O(τ7/4).\displaystyle(-24\tau)^{1/4}-\frac{1}{120}(-24\tau)^{3/4}+\frac{11}{67200}(-24\tau)^{5/4}+O(\tau^{7/4})\,.

Deform the ξ:(,+)\xi:(-\infty,+\infty) integration contour as in the right plot of Fig. 2 along the steepest descent curve Im(h(ξ))=0\mbox{Im}(h(\xi))=0. Denote this curve y0(x)y_{0}(x), where y0(x)y_{0}(x) is the solution satisfying 0<y<π0<y<\pi of sinhxsiny=(πy)x\sinh x\sin y=(\pi-y)x. As in the previous cases, it is sufficient to evaluate only the [S,)[S,\infty) integral, which has the same form as (31).

The expansion of the integrand starts with an inverse quadratic term

(40) sinhξh(ξ)=coshS(ξS)+13!h′′′′(S)(ξS)3+=6(ξS)2+O((ξS)3)\frac{\sinh\xi}{h^{\prime}(\xi)}=\frac{\cosh S(\xi-S)+\cdots}{\frac{1}{3!}h^{\prime\prime\prime\prime}(S)(\xi-S)^{3}+\cdots}=\frac{6}{(\xi-S)^{2}}+O((\xi-S)^{3})

Substituting the derivatives h(k)(S)h^{(k)}(S) gives an explicit expression for the expansion in z:=ξSz:=\xi-S

(41) sinhξh(ξ)=z13!z315!z5O(z7)13!z315!z5O(z7)=sinhzsinhzz.\frac{\sinh\xi}{h^{\prime}(\xi)}=\frac{-z-\frac{1}{3!}z^{3}-\frac{1}{5!}z^{5}-O(z^{7})}{-\frac{1}{3!}z^{3}-\frac{1}{5!}z^{5}-O(z^{7})}=\frac{\sinh z}{\sinh z-z}\,.

Substituting the expansion (39) into (40) gives an expansion of the form

(42) sinhξh(ξ)=321τ+45+13532τ+O(τ)\frac{\sinh\xi}{h^{\prime}(\xi)}=\sqrt{\frac{3}{2}}\frac{1}{\sqrt{-\tau}}+\frac{4}{5}+\frac{1}{35}\sqrt{\frac{3}{2}}\sqrt{-\tau}+O(\tau)

Along the integration path the square root is defined with a negative imaginary part τ=iτ\sqrt{-\tau}=-i\sqrt{\tau}, which gives

(43) Imsinhξh(ξ)=321τ13532τ+O(τ3/2).\mbox{Im}\frac{\sinh\xi}{h^{\prime}(\xi)}=\sqrt{\frac{3}{2}}\frac{1}{\sqrt{\tau}}-\frac{1}{35}\sqrt{\frac{3}{2}}\sqrt{\tau}+O(\tau^{3/2})\,.

Thus the expansion is in powers of τ\tau, rather than fractional powers of τ\tau.

Application of Watson’s lemma gives the expansion of the integral

(44) Im 0e1tτsinhξh(ξ)𝑑τ=e1th(S)32πt(1170t+O(t2))\mbox{Im }\int_{0}^{\infty}e^{-\frac{1}{t}\tau}\frac{\sinh\xi}{h^{\prime}(\xi)}d\tau=e^{-\frac{1}{t}h(S)}\sqrt{\frac{3}{2}\pi t}\left(1-\frac{1}{70}t+O(t^{2})\right)

with h(S)=12π21h(S)=\frac{1}{2}\pi^{2}-1. As in the previous cases, the expansion is in integer powers of tt. The leading term coincides with the ρ1\rho\to 1 limit of the above two cases. The O(t)O(t) subleading term reproduces the first term in Eq. (121) in the Appendix.

The higher order terms O(tj)O(t^{j}) with j>2j>2 for case (iii) can be easily computed.

Remark 2.

The expansion for ρ=1\rho=1 takes a particularly simple form. This is an alternating series, and the first three terms are

(45) θ(1/t,t)=32πte1t(1170t+749,0331,034,880,000t2+O(t3))\theta(1/t,t)=\frac{\sqrt{3}}{2\pi t}e^{\frac{1}{t}}\left(1-\frac{1}{70}t+\frac{749,033}{1,034,880,000}t^{2}+O(t^{3})\right)

2.1. Error estimates

We examine here the error of the approximation for θ(r,t)\theta(r,t) obtained by keeping only the leading asymptotic term in Proposition 1, and give an explicit error bound (53), backed by the results of a numerical study. The bound is uniform in ρ\rho.

Define the remainder of the asymptotic expansion of Proposition 1 as

(46) θ(ρ/t,t)=12πtG(ρ)e1t[F(ρ)π22](1+ϑ(t,ρ)).\theta(\rho/t,t)=\frac{1}{2\pi t}G(\rho)e^{\frac{1}{t}[F(\rho)-\frac{\pi^{2}}{2}]}\left(1+\vartheta(t,\rho)\right)\,.

First, we note that for all cases (i) 0<ρ<10<\rho<1 (ii) ρ>1\rho>1 and (iii) ρ=1\rho=1, the integrand in the application of Watson’s lemma has a similar expansion

(47) Img(ξ)=Img0(ρ)τ+Img2(ρ)τ+O(τ3/2)\mbox{Im}g(\xi)=\frac{\mbox{Im}g_{0}(\rho)}{\sqrt{\tau}}+\mbox{Im}g_{2}(\rho)\sqrt{\tau}+O(\tau^{3/2})

see (28), (35) and (43), respectively. The coefficient of the leading order term is

(48) Img0(ρ)={sinhx12(ρcoshx11),0<ρ<132,ρ=1siny12(1+ρcosy1),ρ>1\mbox{Im}g_{0}(\rho)=\left\{\begin{array}[]{cc}\frac{\sinh x_{1}}{\sqrt{2(\rho\cosh x_{1}-1)}}&\,,0<\rho<1\\ \sqrt{\frac{3}{2}}&\,,\rho=1\\ \frac{\sin y_{1}}{\sqrt{2(1+\rho\cos y_{1})}}&\,,\rho>1\\ \end{array}\right.

We can obtain a bound on the remainder function ϑ(t,ρ)\vartheta(t,\rho) by examining the approximation error of the integrand Im g(ξ)\mbox{Im }g(\xi) by the first term in its series expansion (47). Denoting ξ=x+iy\xi=x+iy we have

(49) Im g(ξ)=xcoshxsiny+(πy)cosysinhx(ρcoshxsinyπ+y)2+(x+ρcosysinhx)2.\mbox{Im }g(\xi)=\frac{x\cosh x\sin y+(\pi-y)\cos y\sinh x}{(\rho\cosh x\sin y-\pi+y)^{2}+(x+\rho\cos y\sinh x)^{2}}\,.

Define

(50) Img(ξ(τ)):=Img0(ρ)τ(1+δ(τ,ρ))\mbox{Im}\,g(\xi(\tau)):=\frac{\mbox{Im}g_{0}(\rho)}{\sqrt{\tau}}(1+\delta(\tau,\rho))

Numerical testing shows the following properties of the remainder function δ(τ,ρ)\delta(\tau,\rho):

i) δ(τ,ρ)0\delta(\tau,\rho)\leq 0 is negative

ii) δ(τ,ρ)1\delta(\tau,\rho)\to-1 as τ\tau\to\infty. This follows by expanding Img(ξ(τ))Img(\xi(\tau)) as xx\to\infty along the steepest descent curves y(x)y(x). For ρ=1\rho=1 this curve has the asymptotics y(x)2πxexy(x)\simeq 2\pi xe^{-x}.

Expressed in terms of τ\tau we get

(51) Img(ξ(τ))=πτ+O(τ2),τ.Img(\xi(\tau))=\frac{\pi}{\tau}+O(\tau^{-2})\,,\quad\tau\to\infty\,.

iii) |δ(τ,ρ)|1|\delta(\tau,\rho)|\leq 1.

iv) it is bounded in absolute value as

(52) |δ(τ,ρ)||g~2(ρ)|τ135τ,τ0|\delta(\tau,\rho)|\leq|\tilde{g}_{2}(\rho)|\tau\leq\frac{1}{35}\tau\,,\quad\tau\geq 0

for all ρ>0\rho>0, where g~2(ρ):=Img2(ρ)Img0(ρ)\tilde{g}_{2}(\rho):=\frac{Img_{2}(\rho)}{Img_{0}(\rho)}. An explicit result for g~2(ρ)\tilde{g}_{2}(\rho) is given in Eq. (120) in the Appendix, and its plot is shown in Fig. 5. Numerical study shows that |g~2(ρ)||\tilde{g}_{2}(\rho)| reaches a maximum value at ρ=1\rho=1, equal to 135\frac{1}{35}. This gives a uniform bound, valid for all ρ>0\rho>0.

We can transfer these results to a bound on the error of the application of Watson’s lemma.

Remark 3.

The error term ϑ(t,ρ)\vartheta(t,\rho) in (46) is bounded from above as

(53) |ϑ(t,ρ)|12|Img2(ρ)||Img0(ρ)|t170t,|\vartheta(t,\rho)|\leq\frac{1}{2}\cdot\frac{|Img_{2}(\rho)|}{|Img_{0}(\rho)|}t\leq\frac{1}{70}t\,,

with g0(ρ),g2(ρ)g_{0}(\rho),g_{2}(\rho) the coefficients in the series expansion (47). The ratio in this bound is denoted as |Img2(ρ)||Img0(ρ)|=g~2(ρ)\frac{|Img_{2}(\rho)|}{|Img_{0}(\rho)|}=\tilde{g}_{2}(\rho) where g~2(ρ)\tilde{g}_{2}(\rho) is given in explicit form in Eq. (120) in the Appendix. It reaches its maximum (in absolute value) at ρ=1\rho=1 with g~2(1)=135\tilde{g}_{2}(1)=-\frac{1}{35}.

Using the property (iii) |δ(τ,ρ)|1|\delta(\tau,\rho)|\leq 1, this bound can be improved as

(54) |ϑ(t,ρ)|min(170t,1).|\vartheta(t,\rho)|\leq\min\left(\frac{1}{70}t,1\right)\,.
Proof.

The approximation error introduced by keeping only the leading order term in the application of Watson’s lemma is estimated as

(55) 0e1tτImg(τ)𝑑τ=0e1tτg0(ρ)τ(1+δ(τ,ρ))𝑑τ:=Img0(ρ)πt(1+ϑ(t,ρ))\int_{0}^{\infty}e^{-\frac{1}{t}\tau}Img(\tau)d\tau=\int_{0}^{\infty}e^{-\frac{1}{t}\tau}\frac{g_{0}(\rho)}{\sqrt{\tau}}(1+\delta(\tau,\rho))d\tau:=Img_{0}(\rho)\sqrt{\pi t}(1+\vartheta(t,\rho))

In the second equality we substituted (50). Using the bound (52) gives the quoted result. ∎

2.2. Properties of the functions F(ρ)F(\rho) and G(ρ)G(\rho)

Let us examine in more detail the properties of the functions F(ρ)F(\rho) and G(ρ)G(\rho) appearing in Proposition 1. We start by studying the behavior of the solutions of the equations (14) and (15) for x1,y1x_{1},y_{1} respectively. For 0<ρ10<\rho\leq 1 the solution of (14) approaches x10x_{1}\to 0 as ρ1\rho\to 1 and it increases to infinity as ρ0\rho\to 0. For ρ1\rho\geq 1, the solution of (15) starts at y1=πy_{1}=\pi for ρ1\rho\to 1, and decreases to zero as ρ\rho\to\infty.

The derivative F(ρ)F^{\prime}(\rho) can be computed exactly

(58) F(ρ)={coshx1,0<ρ<1cosy1,ρ>1\displaystyle F^{\prime}(\rho)=\left\{\begin{array}[]{cc}-\cosh x_{1}\,,&0<\rho<1\\ \cos y_{1}\,,&\rho>1\\ \end{array}\right.

This shows that the minimum of F(ρ)F(\rho) is reached for that value of ρ>1\rho>1 for which y1=π2y_{1}=\frac{\pi}{2}. From (15) it follows that this is reached at ρ=π2\rho=\frac{\pi}{2}, and F(π/2)=3π28F(\pi/2)=\frac{3\pi^{2}}{8}.

We can obtain also the asymptotics of F(ρ)F(\rho) for very small and very large arguments.

Proposition 4.

i) As ρ0\rho\to 0 the function F(ρ)F(\rho) has the asymptotics

(59) F(ρ)=12L2+Llog(2L)L+log2(2L)+π22+o(1)F(\rho)=\frac{1}{2}L^{2}+L\log(2L)-L+\log^{2}(2L)+\frac{\pi^{2}}{2}+o(1)

with L=log(1/ρ)L=\log(1/\rho).

ii) As ρ\rho\to\infty the function F(ρ)F(\rho) has the asymptotics

(60) F(ρ)=ρ+π22(1+ρ)+O(ρ3).F(\rho)=\rho+\frac{\pi^{2}}{2(1+\rho)}+O(\rho^{-3})\,.

iii) Around ρ=1\rho=1, the function F(ρ)F(\rho) has the expansion

F(ρ)\displaystyle F(\rho) =\displaystyle= π221(ρ1)+32(ρ1)265(ρ1)3+351350(ρ1)4\displaystyle\frac{\pi^{2}}{2}-1-(\rho-1)+\frac{3}{2}(\rho-1)^{2}-\frac{6}{5}(\rho-1)^{3}+\frac{351}{350}(\rho-1)^{4}
108125(ρ1)5+372903490000(ρ1)6+O((ρ1)7).\displaystyle-\frac{108}{125}(\rho-1)^{5}+\frac{372903}{490000}(\rho-1)^{6}+O((\rho-1)^{7})\,.
Proof.

i) As ρ0\rho\to 0, the solution of the equation (14) approaches x1x_{1}\to\infty as

(62) x1=L+log(2L)+o(1)x_{1}=L+\log(2L)+o(1)

with L=log(1/ρ)L=\log(1/\rho). This follows from writing Eq. (14) as

(63) 1ρ=sinhx1x1=ex12x1(1e2x1)\frac{1}{\rho}=\frac{\sinh x_{1}}{x_{1}}=\frac{e^{x_{1}}}{2x_{1}}(1-e^{-2x_{1}})

Taking the logs of both sides gives

(64) x1=L+log(2x1)log(1e2x1).x_{1}=L+\log(2x_{1})-\log(1-e^{-2x_{1}})\,.

This can be iterated starting with the zero-th approximation x1(0)=Lx_{1}^{(0)}=L, which gives (62).

Eliminating ρ\rho in terms of x1x_{1} using (14) gives

(65) F(ρ)=12x12x1tanhx1+π22F(\rho)=\frac{1}{2}x_{1}^{2}-\frac{x_{1}}{\tanh x_{1}}+\frac{\pi^{2}}{2}

Substituting (62) into this expression gives the quoted result.

ii) As ρ\rho\to\infty, the solution of the equation (15) approaches y10y_{1}\to 0. This equation is approximated as

(66) (1+ρ)y1+16ρy13+ρO(y15)=π(1+\rho)y_{1}+\frac{1}{6}\rho y_{1}^{3}+\rho O(y_{1}^{5})=\pi

which is inverted as

(67) y1=π1+ρπ36ρ(1+ρ)4+O(ρ5)y_{1}=\frac{\pi}{1+\rho}-\frac{\pi^{3}}{6}\frac{\rho}{(1+\rho)^{4}}+O(\rho^{-5})

Substituting into F(ρ)=12y12+ρcosy1+πy1F(\rho)=-\frac{1}{2}y_{1}^{2}+\rho\cos y_{1}+\pi y_{1} gives the result quoted above.

iii) Consider first the case 0<ρ10<\rho\leq 1. As ρ1\rho\to 1, we have x10x_{1}\to 0. From (14) we get an expansion x12=6(ρ1)+215(ρ1)2564175(ρ1)3+O((ρ1)4)x_{1}^{2}=-6(\rho-1)+\frac{21}{5}(\rho-1)^{2}-\frac{564}{175}(\rho-1)^{3}+O((\rho-1)^{4}). Substituting into (65) gives the result (4). The same result result is obtained for the ρ1\rho\geq 1 case, when ρ=1\rho=1 is approached from above.

We give next the asymptotics of G(ρ)G(\rho).

Proposition 5.

i) As ρ0\rho\to 0 the asymptotics of G(ρ)G(\rho) is

(68) G(ρ)=Lρ2L+log2L+12L+o(1),L:=log(1/ρ).G(\rho)=\sqrt{L}-\rho^{2}\sqrt{L}+\frac{\log 2L+1}{2\sqrt{L}}+o(1)\,,\quad L:=\log(1/\rho)\,.

ii) As ρ\rho\to\infty we have

(69) G(ρ)=πρ(1+ρ)3/2+O(ρ5/2).G(\rho)=\frac{\pi\rho}{(1+\rho)^{3/2}}+O(\rho^{-5/2})\,.

iii) Around ρ=1\rho=1 the function G(ρ)G(\rho) has the following expansion

(70) G(ρ)\displaystyle G(\rho) =\displaystyle= 3{1+15(1ρ1)435(1ρ1)2+225(1ρ1)3+O((1ρ1)4)}.\displaystyle\sqrt{3}\left\{1+\frac{1}{5}\left(\frac{1}{\rho}-1\right)-\frac{4}{35}\left(\frac{1}{\rho}-1\right)^{2}+\frac{2}{25}\left(\frac{1}{\rho}-1\right)^{3}+O(\left(\frac{1}{\rho}-1\right)^{4})\right\}\,.
Proof.

i) The asymptotic expansion of x1x_{1}\to\infty for ρ0\rho\to 0 was already obtained in (62). Substitute this into

(71) G(ρ)=x1x1tanhx11G(\rho)=\frac{x_{1}}{\sqrt{\frac{x_{1}}{\tanh x_{1}}-1}}

which follows from (13) by eliminating ρ\rho using (14). Expanding the result gives the result quoted.

ii) Eliminating ρ\rho from the expression (13) for G(ρ)G(\rho) for ρ1\rho\geq 1 gives

(72) G(ρ)=πy11+πy1tany1G(\rho)=\frac{\pi-y_{1}}{\sqrt{1+\frac{\pi-y_{1}}{\tan y_{1}}}}

Using here the expansion of y1y_{1} from (67) gives the result quoted.

iii) The proof is similar to that of point (iii) in Proposition 4. ∎

3. Numerical tests

The leading term of the expansion in Proposition 1 can be considered as an approximation for θ(r,t)\theta(r,t) for all t0t\geq 0, by substituting ρ=rt\rho=rt. Consider the approximation for θ(r,t)\theta(r,t) defined as

(73) θ^(r,t):=12πtG(rt)e1t(F(rt)π22).\hat{\theta}(r,t):=\frac{1}{2\pi t}G(rt)e^{-\frac{1}{t}(F(rt)-\frac{\pi^{2}}{2})}\,.

We would like to compare this approximation with the leading asymptotic expansion of [12], which is obtained by taking the limit t0t\to 0 at fixed rr. This is given by Theorem 1 in [12]

(74) θG(r,t)=eπu0(t)logu0(t)22ρetu0(t)+2u0(t)\theta_{G}(r,t)=\frac{\sqrt{e}}{\pi}\sqrt{\frac{u_{0}(t)}{\log u_{0}(t)-2-2\rho}}e^{-tu_{0}(t)+\sqrt{2u_{0}(t)}}

where u0(t)u_{0}(t) is the solution of the equation

(75) t=logu22uρ2u+14u,ρ:=logr22.t=\frac{\log u}{2\sqrt{2u}}-\frac{\rho}{\sqrt{2u}}+\frac{1}{4u}\,,\quad\rho:=\log\frac{r}{2\sqrt{2}}\,.

The correction to Eq. (74) is of order 1+O(tlog2(1/t))1+O(\sqrt{t}\log^{2}(1/t)).

Table 1 shows the numerical evaluation of θ^(r,t)\hat{\theta}(r,t) for r=0.5r=0.5 and several values of tt, comparing also with the small tt expansion θG(r,t)\theta_{G}(r,t) in Eq. (74). They agree well for sufficiently small tt but start to diverge for larger tt. In the last column of Table 1 we show also the result of a direct numerical evaluation of θ(r,t)\theta(r,t) by numerical integration in Mathematica.

Table 1. Numerical evaluation of θ(0.5,t)\theta(0.5,t) using the asymptotic expansion of Proposition 1 θ^(0.5,t)\hat{\theta}(0.5,t) and the Gerhold approximation θG(0.5,t)\theta_{G}(0.5,t) given in (74). The last column shows the result of a direct numerical evaluation using numerical integration in Mathematica.
tt ρ\rho x1x_{1} F(ρ)F(\rho) θ^(0.5,t)\hat{\theta}(0.5,t) u0(t)u_{0}(t) θG(0.5,t)\theta_{G}(0.5,t) θnum(0.5,t)\theta_{\rm num}(0.5,t)
0.1 0.05 5.3697 13.9816 2.09810392.098\cdot 10^{-39} 1447.8 2.10110392.101\cdot 10^{-39} -
0.2 0.1 4.4999 10.5584 1.17610121.176\cdot 10^{-12} 256.3 1.18110121.181\cdot 10^{-12} 1.17310121.173\cdot 10^{-12}
0.3 0.15 3.9692 8.84 2.7131062.713\cdot 10^{-6} 89.713 2.7381062.738\cdot 10^{-6} 2.7041062.704\cdot 10^{-6}
0.5 0.25 3.2638 6.9876 0.0114 22.69 0.0116 0.0113
1.0 0.5 2.1773 5.0712 0.2722 3.1345 0.3062 0.2685
1.5 0.75 1.3512 4.3023 0.2960 0.9531 0.4097 0.2900
2.0 1.0 0.0 3.9348 0.2300 0.4271 0.4690 0.2213
tt ρ\rho y1y_{1} F(ρ)F(\rho) θ^(0.5,t)\hat{\theta}(0.5,t) u0(t)u_{0}(t) θG(0.5,t)\theta_{G}(0.5,t) θnum(0.5,t)\theta_{\rm num}(0.5,t)
2.5 1.25 2.0105 3.7630 0.1682 0.2430 1.2541 0.1628
3.0 1.5 1.6458 3.7037 0.127 0.1607 - 0.1222
10.0 5.0 0.5459 5.8393 0.0164 0.0234 - 0.0151

Figure 3 compares the approximation θ^(r,t)\hat{\theta}(r,t) (black curves) against θG(r,t)\theta_{G}(r,t) (blue curves) and direct numerical integration (red curves). We note the same pattern of good agreement between θ^(r,t)\hat{\theta}(r,t) and θG(r,t)\theta_{G}(r,t) at small tt, and increasing discrepancy for larger tt, which explodes to infinity for sufficiently large tt. The reason for this explosive behavior is the fact that the denominator in Eq. (74) approaches zero as tt approaches a certain maximum value. For tt larger than this value the denominator becomes negative and the approximation ceases to exist.

This maximum tt value is given by the inequality u0(t)<e22log(r/(22))=8r2e2u_{0}(t)<e^{-2-2\log(r/(2\sqrt{2}))}=\frac{8}{r^{2}e^{2}}. For r=0.5r=0.5 this condition imposes the upper bound t<tmax=2.5538t<t_{\rm max}=2.5538.

Refer to caption
Refer to caption
Refer to caption
Figure 3. Plot of θ^(r,t)\hat{\theta}(r,t) vs tt from the asymptotic expansion of Proposition 1 (black) and from the Theorem 1 of Gerhold [12] (blue). The red curves show the results of direct numerical integration of θ(r,t)\theta(r,t). The three panels correspond to the three values of r=0.5,1.0,1.5r=0.5,1.0,1.5.
Refer to caption
Figure 4. Plot of θ^(r,t)\hat{\theta}(r,t) vs tt defined in (73) giving the leading asymptotic result from Proposition 1 for three values of r=0.5,1.0,1.5r=0.5,1.0,1.5 (solid, dashed, dotted). The widths of the curves reflect the error bound (53).

We show in Figure 4 plots of θ^(r,t)\hat{\theta}(r,t) vs tt at r=0.5,1.0,1.5r=0.5,1.0,1.5. The vertical scale of the plot is chosen as in Figure 1 (left) of Bernhart and Mai [4], which shows the plots of θ(r,t)\theta(r,t) for the same parameters, obtained using a precise numerical inversion of the Laplace transform of θ(r,t)\theta(r,t). The shapes of the curves are very similar with those shown in [4].

3.1. Asymptotics for t0t\to 0 and tt\to\infty of the approximation θ^(r,t)\hat{\theta}(r,t)

We study in this Section the asymptotics of the approximation θ^(r,t)\hat{\theta}(r,t) defined in (73) for t0t\to 0 and tt\to\infty, and compare with the exact asymptotics of θ(r,t)\theta(r,t) in these limits obtained by Barrieu et al [3] and Gerhold [12].

As expected, the t0t\to 0 asymptotics matches precisely the exact asymptotics of θ(r,t)\theta(r,t). Although the approximation θ^(r,t)\hat{\theta}(r,t) was derived in the small tt limit, it is somewhat surprising that it matches also the correct asymptotics of θ(r,t)\theta(r,t) for tt\to\infty, up to a multiplicative coefficient, which however becomes exact as rr\to\infty.

Small tt asymptotics t0t\to 0. The leading asymptotics of θ(r,t)\theta(r,t) as t0t\to 0 was obtained in Barrieu et al [3]

(76) θ(r,t)e12tlog2t.\theta(r,t)\sim e^{-\frac{1}{2t}\log^{2}t}\,.

This was refined further by Gerhold as in Eq. (74). Using the small tt approximation u0(t)=12t2log2(1/t)u_{0}(t)=\frac{1}{2t^{2}}\log^{2}(1/t) (see Eq. (6) in Gerhold’s paper), the improved expansion (74) gives

(77) θG(r,t)log(1/t)t12loglog(1/t)+2log(2/t)e12tlog2(1/t).\theta_{G}(r,t)\sim\frac{\log(1/t)}{t}\frac{1}{\sqrt{2\log\log(1/t)+2\log(2/t)}}e^{-\frac{1}{2t}\log^{2}(1/t)}\,.

The t0t\to 0 expansion of θ^(r,t)\hat{\theta}(r,t) can be obtained using the ρ0\rho\to 0 asymptotics of F(ρ),G(ρ)F(\rho),G(\rho) in points (i) of the Propositions 4 and 5, respectively. This gives

(78) θ^(r,t)1tlog1/(rt)e12tlog2(rt)1tlog(1/(rt))log(2log(1/(rt))+1tlog(1/(rt)),t0\hat{\theta}(r,t)\sim\frac{1}{t}\sqrt{\log 1/(rt)}e^{-\frac{1}{2t}\log^{2}(rt)-\frac{1}{t}\log(1/(rt))\log(2\log(1/(rt))+\frac{1}{t}\log(1/(rt))}\,,\quad t\to 0

The exponential factor agrees with the asymptotics of the exact result. Also the leading dependence of the multiplying factor reproduces the improved expansion following from [12].

Large tt asymptotics tt\to\infty. The tt\to\infty asymptotics of θ(r,t)\theta(r,t) is given in Remark 3 in Barrieu et al [3] as

(79) θ(r,t)cr1t3/2,cr=12πK0(r).\theta(r,t)\sim c_{r}\frac{1}{t^{3/2}}\,,\quad c_{r}=\frac{1}{\sqrt{2\pi}}K_{0}(r)\,.

The large tt asymptotics of θ^(r,t)\hat{\theta}(r,t) is related to the ρ\rho\to\infty asymptotics of F(ρ),G(ρ)F(\rho),G(\rho). As ρ\rho\to\infty we have F(ρ)ρF(\rho)\sim\rho from Prop. 4 point (ii) and G(ρ)πρ(1+ρ)3/2G(\rho)\sim\frac{\pi\rho}{(1+\rho)^{3/2}} from Prop. 5 point (ii). Substituting into (73) and keeping only the leading contributions as tt\to\infty gives

(80) θ^(r,t)r2(1+rt)3/2er12rt3/2er.\hat{\theta}(r,t)\sim\frac{r}{2(1+rt)^{3/2}}e^{-r}\sim\frac{1}{2\sqrt{r}}t^{-3/2}e^{-r}\,.

The tt dependence has the same form as the exact asymptotics (79).

Let us examine also the coefficient of t3/2t^{-3/2} in (80) and compare with the exact result for crc_{r} in (79). The leading asymptotics of the K0(x)K_{0}(x) function for xx\to\infty is K(x)=exπ2x(1+O(1/x))K(x)=e^{-x}\sqrt{\frac{\pi}{2x}}(1+O(1/x)). The exact coefficient becomes, for rr\to\infty

(81) cr=121rer+O(1/r)c_{r}=\frac{1}{2}\frac{1}{\sqrt{r}}e^{-r}+O(1/r)

The leading term in this expansion matches precisely the coefficient obtained from θ^(r,t)\hat{\theta}(r,t). We conclude that the right tail asymptotics of θ^(r,t)\hat{\theta}(r,t) approaches the same form as the exact asymptotics in the limit rr\to\infty.

4. Application: The asymptotic distribution of 1tAt(μ)\frac{1}{t}A_{t}^{(\mu)}

As an application of the result of Proposition 1, we give here the leading t0t\to 0 asymptotics of the distribution of the time average of the gBM 1tAt(μ)\frac{1}{t}A_{t}^{(\mu)}.

The distribution of 1tAt(μ)\frac{1}{t}A_{t}^{(\mu)} can be obtained from Yor’s formula (3) by integration over xx. Introducing a new variable u=atu=at this is expressed as the integral

(82) (1tAt(μ)da)=xeμx12μ2texp(1+e2x2at)θ(exat,t)dadxa.\mathbb{P}\left(\frac{1}{t}A_{t}^{(\mu)}\in da\right)=\int_{x\in\mathbb{R}}e^{\mu x-\frac{1}{2}\mu^{2}t}\exp\left(-\frac{1+e^{2x}}{2at}\right)\theta\left(\frac{e^{x}}{at},t\right)\frac{dadx}{a}\,.

Changing the xx integration variable to x=log(aρ)x=\log(a\rho), the range of integration for x:(,)x:(-\infty,\infty) is mapped to ρ:(0,)\rho:(0,\infty).

(83) (1tAt(μ)da)=e12μ2tdaa0(aρ)μe1+a2ρ22atθ(ρt,t)dρρ.\displaystyle\mathbb{P}\left(\frac{1}{t}A_{t}^{(\mu)}\in da\right)=e^{-\frac{1}{2}\mu^{2}t}\frac{da}{a}\int_{0}^{\infty}(a\rho)^{\mu}e^{-\frac{1+a^{2}\rho^{2}}{2at}}\theta\left(\frac{\rho}{t},t\right)\frac{d\rho}{\rho}\,.

We would like to use here the asymptotic expansion for θ(ρ/t,t)\theta(\rho/t,t) from Proposition 1. Integrating term by term the asymptotic expansion requires a uniform bound on the approximation error. Such a bound was given in (53) based on numerical evidence, which is furthermore uniform in ρ\rho. Using this result one can transfer the asymptotic expansion of Proposition 1 to an asymptotic result for the density of the time-average of the gBM. Define this density as

(1tAt(μ)da)=f(a,t)daa.\mathbb{P}\left(\frac{1}{t}A_{t}^{(\mu)}\in da\right)=f(a,t)\frac{da}{a}\,.
Proposition 6.

Assume the error bound (53). Then we have

(84) f(a,t)=12πte12μ2t0(aρ)μe1tH(ρ)G(ρ)dρρ(1+ε(a,t))f(a,t)=\frac{1}{2\pi t}e^{-\frac{1}{2}\mu^{2}t}\int_{0}^{\infty}(a\rho)^{\mu}e^{-\frac{1}{t}H(\rho)}G(\rho)\frac{d\rho}{\rho}\cdot(1+\varepsilon(a,t))

with

(85) H(ρ):=1+a2ρ22aπ22+F(ρ).H(\rho):=\frac{1+a^{2}\rho^{2}}{2a}-\frac{\pi^{2}}{2}+F(\rho)\,.

The error term is bounded as |ε(a,t)|170t|\varepsilon(a,t)|\leq\frac{1}{70}t.

Furthermore, the density f(a,t)f(a,t) has the asymptotic form in the small-maturity limit

(86) f(a,t)12πtg(a,μ)e1tJ(a),t0,f(a,t)\sim\frac{1}{\sqrt{2\pi t}}g(a,\mu)e^{-\frac{1}{t}J(a)}\,,\quad t\to 0\,,

with J(a)infρ0H(ρ)=H(ρ)J(a)\equiv\inf_{\rho\geq 0}H(\rho)=H(\rho_{*}) and

(87) g(a,μ):=(aρ)μG(ρ)1ρH′′(ρ),ρ=arginfρH(ρ).g(a,\mu):=(a\rho_{*})^{\mu}G(\rho_{*})\frac{1}{\rho_{*}\sqrt{H^{\prime\prime}(\rho_{*})}}\,,\quad\rho_{*}=\arg\inf_{\rho}H(\rho)\,.
Proof.

The function θ(ρ,t)\theta(\rho,t) can be expressed as the leading order asymptotic term in Proposition 1

(88) θ(ρ/t,t)=12πte1t(F(ρ)π22)G(ρ)(1+ϑ(ρ,t)),\theta(\rho/t,t)=\frac{1}{2\pi t}e^{-\frac{1}{t}(F(\rho)-\frac{\pi^{2}}{2})}G(\rho)\Big{(}1+\vartheta(\rho,t)\Big{)}\,,

where the error term ϑ(ρ,t)\vartheta(\rho,t) is bounded as in (53).

Substituting into the integral in (83) gives

(89) 0(aρ)μe1+a2ρ22atθ(ρt,t)dρρ=12πt0(aρ)μe1tH(ρ)G(ρ)dρρ+Iε(a,t)\displaystyle\int_{0}^{\infty}(a\rho)^{\mu}e^{-\frac{1+a^{2}\rho^{2}}{2at}}\theta\left(\frac{\rho}{t},t\right)\frac{d\rho}{\rho}=\frac{1}{2\pi t}\int_{0}^{\infty}(a\rho)^{\mu}e^{-\frac{1}{t}H(\rho)}G(\rho)\frac{d\rho}{\rho}+I_{\varepsilon}(a,t)

with ρ=arginfρH(ρ)\rho_{*}=\arg\inf_{\rho}H(\rho).

The term Iε(a,t)I_{\varepsilon}(a,t) is bounded using the error bound (53)

(90) |Iε(a,t)|0(aρ)μe1+a2ρ22atθ^(ρ,t)|ϑ(ρ,t)|dρρ170t0(aρ)μe1+a2ρ22atθ^(ρ,t)dρρ\displaystyle|I_{\varepsilon}(a,t)|\leq\int_{0}^{\infty}(a\rho)^{\mu}e^{-\frac{1+a^{2}\rho^{2}}{2at}}\hat{\theta}(\rho,t)|\vartheta(\rho,t)|\frac{d\rho}{\rho}\leq\frac{1}{70}t\int_{0}^{\infty}(a\rho)^{\mu}e^{-\frac{1+a^{2}\rho^{2}}{2at}}\hat{\theta}(\rho,t)\frac{d\rho}{\rho}

which gives |ε(t)|170t|\varepsilon(t)|\leq\frac{1}{70}t, as stated.

The asymptotic result (86) follows by application of the Laplace asymptotic expansion [11] to the integral (84). ∎

4.1. Properties of J(a)J(a)

In this section we give a more explicit result for the exponential factor J(a)J(a) in the leading asymptotic density (84), and study its properties.

Proposition 7.

The function J(a)J(a) is given by:

i) for a1a\geq 1 we have

(91) J(a)=12a12acosh2x1+12x12=12x1212x1tanhx1\displaystyle J(a)=\frac{1}{2a}-\frac{1}{2a}\cosh^{2}x_{1}+\frac{1}{2}x_{1}^{2}=\frac{1}{2}x_{1}^{2}-\frac{1}{2}x_{1}\tanh x_{1}

where x1x_{1} is the solution of the equation

(92) sinh2x12x1=a.\frac{\sinh 2x_{1}}{2x_{1}}=a\,.

This case corresponds to 0<ρ10<\rho_{*}\leq 1.

ii) for 0<a10<a\leq 1 we have

(93) J(a)=12(y1π)tan(y1π)12(y1π)2J(a)=\frac{1}{2}(y_{1}-\pi)\tan(y_{1}-\pi)-\frac{1}{2}(y_{1}-\pi)^{2}

where y1y_{1} is the solution of the equation

(94) y11asiny1cosy1=π.y_{1}-\frac{1}{a}\sin y_{1}\cos y_{1}=\pi\,.

This case corresponds to 1ρ<π21\leq\rho_{*}<\frac{\pi}{2}.

Proof.

We start with the expression

J(a)\displaystyle J(a) =\displaystyle= infρ0(1+a2ρ22aπ22+F(ρ))\displaystyle\inf_{\rho\geq 0}\left(\frac{1+a^{2}\rho^{2}}{2a}-\frac{\pi^{2}}{2}+F(\rho)\right)
=\displaystyle= 12a+a2ρ2π22+F(ρ)\displaystyle\frac{1}{2a}+\frac{a}{2}\rho_{*}^{2}-\frac{\pi^{2}}{2}+F(\rho_{*})

where ρ\rho_{*} is the solution of the equation

(96) F(ρ)+aρ=0.F^{\prime}(\rho)+a\rho=0\,.

Using the explicit result for F(ρ)F^{\prime}(\rho) in Eq. (58) we get F(ρ)<0F^{\prime}(\rho)<0 for 0<ρ<π20<\rho<\frac{\pi}{2} and F(ρ)>0F^{\prime}(\rho)>0 for ρ>π2\rho>\frac{\pi}{2}. This implies that the solution of the equation (96) must satisfy ρπ2\rho_{*}\leq\frac{\pi}{2}. As a0a\to 0 the minimizer approaches the upper boundary lima0ρ=π2\lim_{a\to 0}\rho_{*}=\frac{\pi}{2} as F(π2)=0F^{\prime}(\frac{\pi}{2})=0.

The solution of (96) for a=1a=1 is ρ=1\rho_{*}=1, since F(1)=1F^{\prime}(1)=-1, from Prop. 4 point (iii).

The strategy of the proof is to eliminate ρ\rho_{*} using (58) in terms of x1x_{1} and y1y_{1}, respectively. Substituting (58) into (96) gives

(97) 0<ρ<1\displaystyle 0<\rho<1 :\displaystyle: coshx1=aρ\displaystyle\cosh x_{1}=a\rho
(98) ρ>1\displaystyle\rho>1 :\displaystyle: cosy1=aρ.\displaystyle\cos y_{1}=-a\rho\,.

These equations can be used to eliminate ρ\rho_{*} from the expression of F(ρ)F(\rho_{*}). Substituting into (4.1) reproduces the results shown.

The function ρ(a)\rho_{*}(a) has the expansion around a=1a=1

(99) ρ(a)=114(a1)+19160(a1)2151122400(a1)3+O((a1)4).\rho_{*}(a)=1-\frac{1}{4}(a-1)+\frac{19}{160}(a-1)^{2}-\frac{1511}{22400}(a-1)^{3}+O((a-1)^{4})\,.

To prove this expansion, assume first a1a\to 1 from above, and thus ρ<1\rho_{*}<1. From Eq. (97) it follows that

(100) ρ=1acoshx1(a)\rho_{*}=\frac{1}{a}\cosh x_{1}(a)

where x1(a)x_{1}(a) is the solution of sinh2x12x1=a\frac{\sinh 2x_{1}}{2x_{1}}=a. Expanding around a=1a=1 gives x1(a)=32(a1)+x_{1}(a)=\sqrt{\frac{3}{2}(a-1)}+\cdots, and substituting above gives the expansion of ρ(a)\rho_{*}(a). A similar result is obtained by assuming a1a\to 1 from below.

The function J(a)J(a) has the following properties:

i) The function J(a)J(a) vanishes at a=1a=1. The infimum in (4.1) is reached at ρ=1\rho_{*}=1.

ii) For a>1a>1 the infimum is reached at 0<ρ<10<\rho_{*}<1. As aa\to\infty we have limaρ=0\lim_{a\to\infty}\rho_{*}=0.

iii) For 0<a<10<a<1 the infimum is reached at 1<ρ<ρ0=π21<\rho_{*}<\rho_{0}=\frac{\pi}{2}. As a0a\to 0 we have lima0ρ=π/2\lim_{a\to 0}\rho_{*}=\pi/2.

4.2. Comparison with the Large Deviations result

The distributional properties of At(μ)A_{t}^{(\mu)} as t0t\to 0 were studied using probabilistic methods from Large Deviations theory in [2, 21, 23]. Assuming that StS_{t} follows a one-dimensional diffusion dSt=σ(St)StdWt+(rq)StdtdS_{t}=\sigma(S_{t})S_{t}dW_{t}+(r-q)S_{t}dt, it has been shown in [21] that, under weak assumptions on σ()\sigma(\cdot), (Att)\mathbb{P}\left(\frac{A_{t}}{t}\in\cdot\right) satisfies a Large Deviations property as t0t\to 0. This result includes as a limiting case the geometric Brownian motion corresponding to σ(x)=σ\sigma(x)=\sigma.

The main result can be extracted from the proof of Theorem 2 in [21], and can be formulated as follows.

Proposition 8.

Assume St=eσWt+(r12σ2)tS_{t}=e^{\sigma W_{t}+(r-\frac{1}{2}\sigma^{2})t} and At=0tSt𝑑tA_{t}=\int_{0}^{t}S_{t}dt. The time average of StS_{t} satisfies a Large Deviations property in the t0t\to 0 limit

(101) limt0tlog(Att)=1σ2𝒥BS()\lim_{t\to 0}t\log\mathbb{P}\left(\frac{A_{t}}{t}\in\cdot\right)=-\frac{1}{\sigma^{2}}\mathcal{J}_{\rm BS}(\cdot)

with rate function 𝒥BS()\mathcal{J}_{\rm BS}(\cdot) expressed as the solution of a variational problem, which is furthermore solved in closed form in Proposition 12 of [21].

The asymptotic density (84) has the same form as the exponential decay predicted by Large Deviations theory (101). We show here that the result for J(a)J(a) in Proposition 7 is related to the rate function 𝒥BS(a)\mathcal{J}_{\rm BS}(a) derived in [21] as

(102) J(a)=14𝒥BS(a).J(a)=\frac{1}{4}\mathcal{J}_{\rm BS}(a)\,.

The factor of 1/4 is due to the factor of 2 in the exponent of the definition of AtA_{t} in the Yor formula (3), which acts as a volatility σ=2\sigma=2.

The function 𝒥BS(x)\mathcal{J}_{\rm BS}(x) is given by Proposition 12 in [21] which we reproduce here for convenience.

Proposition 9 (Prop. 12 in [21]).

The rate function 𝒥BS(x)\mathcal{J}_{\rm BS}(x) is given by

(105) 𝒥BS(x)={12β2βtanh(β2),x12ξ(tanξξ),x1\displaystyle\mathcal{J}_{\rm BS}(x)=\left\{\begin{array}[]{cc}\frac{1}{2}\beta^{2}-\beta\tanh\left(\frac{\beta}{2}\right)\,,&x\geq 1\\ 2\xi(\tan\xi-\xi)\,,&x\leq 1\\ \end{array}\right.

where β\beta is the solution of the equation

(106) sinhββ=x\frac{\sinh\beta}{\beta}=x

and ξ\xi is the solution in [0,π/2][0,\pi/2] of the equation

(107) sin2ξ2ξ=x.\frac{\sin 2\xi}{2\xi}=x\,.

The relation (102) follows by identifying x1=12βx_{1}=\frac{1}{2}\beta in the first case of Proposition 7, and y1π=ξy_{1}-\pi=\xi in the second case.

The rate function 𝒥BS(a)\mathcal{J}_{\rm BS}(a) for a1a\simeq 1 can be expanded in a Taylor series in loga\log a (see Eq. (36) of [21])

(108) 𝒥BS(a)=32log2a310log3a+1091400log4a+O(log5a).\mathcal{J}_{\rm BS}(a)=\frac{3}{2}\log^{2}a-\frac{3}{10}\log^{3}a+\frac{109}{1400}\log^{4}a+O(\log^{5}a)\,.

This expansion is convenient for numerical evaluation of the rate function. Proposition 13 in [21] gives also asymptotic expansions of 𝒥BS(a)\mathcal{J}_{\rm BS}(a) for a0a\to 0 and aa\to\infty, which can be translated directly into corresponding asymptotic expansions for J(a)J(a).

4.3. Properties of g(a,μ)g(a,\mu)

We study here in more detail the properties of the function g(a,μ)g(a,\mu) defined in (87). This can be put into a more explicit form as follows.

Proposition 10.

The function g(a,μ)g(a,\mu) has the explicit form

(109) g(a,μ)\displaystyle g(a,\mu) =\displaystyle= eμloga+(μ1)logρ(a)G(ρ)1H′′(ρ)=32ec1loga+c2log2a+O(log3a).\displaystyle e^{\mu\log a+(\mu-1)\log\rho_{*}(a)}G(\rho_{*})\frac{1}{\sqrt{H^{\prime\prime}(\rho_{*})}}=\frac{\sqrt{3}}{2}e^{c_{1}\log a+c_{2}\log^{2}a+O(\log^{3}a)}\,.

with

(110) c1=34(μ+1)45\displaystyle c_{1}=\frac{3}{4}(\mu+1)-\frac{4}{5}
(111) c2=380(μ+1)+571400.\displaystyle c_{2}=-\frac{3}{80}(\mu+1)+\frac{57}{1400}\,.
Proof.

The proof follows by combining the expansions around a=1a=1 of the factors in this expression:

i) ρ(a)\rho_{*}(a) is given by (99) which is written in equivalent form as

(112) logρ(a)=14loga380log2a+1350log3a+O(log4a).\log\rho_{*}(a)=-\frac{1}{4}\log a-\frac{3}{80}\log^{2}a+\frac{1}{350}\log^{3}a+O(\log^{4}a)\,.

ii) G(ρ)G(\rho_{*}) is obtained by substituting the expansion of ρ(a)\rho_{*}(a) into the expansion of G(ρ)G(\rho) in powers of ρ1\rho-1 from Proposition 5

G(ρ(a))\displaystyle G(\rho_{*}(a)) =\displaystyle= 3(1+120loga+375600log2a4148000log3a+O(log4a))\displaystyle\sqrt{3}\left(1+\frac{1}{20}\log a+\frac{37}{5600}\log^{2}a-\frac{41}{48000}\log^{3}a+O(\log^{4}a)\right)
=\displaystyle= 3exp(120loga+3560log2a1875log3a+O(log4a))\displaystyle\sqrt{3}\exp\left(\frac{1}{20}\log a+\frac{3}{560}\log^{2}a-\frac{1}{875}\log^{3}a+O(\log^{4}a)\right)

iii) H′′(ρ(a))=F′′(ρ(a))+aH^{\prime\prime}(\rho_{*}(a))=F^{\prime\prime}(\rho_{*}(a))+a can be evaluated using the expansion of F(ρ)F(\rho) from Prop. 4 (iii) which gives

(114) F′′(ρ)=3+95(a1)18175(a1)2+O((a1)3).F^{\prime\prime}(\rho_{*})=3+\frac{9}{5}(a-1)-\frac{18}{175}(a-1)^{2}+O((a-1)^{3})\,.

We get

(115) H′′(ρ(a))=4+145(a1)18175(a1)2+O((a1)3).H^{\prime\prime}(\rho_{*}(a))=4+\frac{14}{5}(a-1)-\frac{18}{175}(a-1)^{2}+O((a-1)^{3})\,.

It is convenient to express this in terms of loga\log a as

(116) H′′(ρ)=4exp(710loga+1111400log2a+O(log3a)).H^{\prime\prime}(\rho_{*})=4\exp\left(\frac{7}{10}\log a+\frac{111}{1400}\log^{2}a+O(\log^{3}a)\right)\,.

Remark 11.

We have g(1,μ)=32g(1,\mu)=\frac{\sqrt{3}}{2} for all μ\mu\in\mathbb{R}. This follows by noting that at a=1a=1 we have ρ(1)=1\rho_{*}(1)=1, and thus the only factors different from 1 are G(1)=3G(1)=\sqrt{3} and H′′(1)=4H^{\prime\prime}(1)=4.

Numerical evaluation of the function g(a,μ)g(a,\mu) shows that for the driftless gBM case μ=1\mu=-1 this function is decreasing in aa, and becomes increasing for sufficiently large and positive μ\mu.

Refer to caption
Figure 5. Plot of the function g~2(ρ)\tilde{g}_{2}(\rho) appearing in the subleading correction to the asymptotic expansion of the Hartman-Watson integral (117).

Acknowledgements

I am grateful to Peter Nándori and Lingjiong Zhu for very useful discussions and comments.

Appendix A Subleading correction

We give here the subleading correction to the asymptotic expansion of the Hartman-Watson integral θ(ρ/t,t)\theta(\rho/t,t) in Proposition 1.

The first two terms of the asymptotic expansion of the Hartman-Watson integral as t0t\to 0 are

(117) θ(ρ/t,t)=12πtG(ρ)e1t(F(ρ)π22)(1+12tg~2(ρ)+O(t2))\theta(\rho/t,t)=\frac{1}{2\pi t}G(\rho)e^{-\frac{1}{t}(F(\rho)-\frac{\pi^{2}}{2})}\left(1+\frac{1}{2}t\tilde{g}_{2}(\rho)+O(t^{2})\right)

where

(120) g~2(ρ)={12+9ρcoshx12ρ2cosh2x1+5ρ212(ρcoshx11)3,0<ρ112+9ρcosy1+2ρ2cos2y15ρ212(1+ρcosy1)3,ρ>1\displaystyle\tilde{g}_{2}(\rho)=\left\{\begin{array}[]{cc}\frac{-12+9\rho\cosh x_{1}-2\rho^{2}\cosh^{2}x_{1}+5\rho^{2}}{12(\rho\cosh x_{1}-1)^{3}}&\,,0<\rho\leq 1\\ \frac{12+9\rho\cos y_{1}+2\rho^{2}\cos^{2}y_{1}-5\rho^{2}}{12(1+\rho\cos y_{1})^{3}}&\,,\rho>1\\ \end{array}\right.

The result follows straighforwardly by keeping the next terms in the expansions of the integrals appearing in the proof of Proposition 1.

The plot of g~2(ρ)\tilde{g}_{2}(\rho) is given in the left panel of Figure 5. It reaches a maximum (in absolute value) at ρ=1\rho=1 where it is equal to g~2(1)=135\tilde{g}_{2}(1)=-\frac{1}{35}.

We list next a few properties of the function g~2(ρ)\tilde{g}_{2}(\rho). This function has the expansion around ρ=1\rho=1

(121) g~2(ρ)=135+14467375(1/ρ1)+O((1/ρ1)2).\tilde{g}_{2}(\rho)=-\frac{1}{35}+\frac{144}{67375}(1/\rho-1)+O((1/\rho-1)^{2})\,.

As ρ\rho\to\infty we get, using the asymptotics of y10y_{1}\to 0 from the proof of Prop. 4(ii),

(122) g~2(ρ)=14ρ+32ρ2+O(ρ3).\tilde{g}_{2}(\rho)=-\frac{1}{4\rho}+\frac{3}{2\rho^{2}}+O(\rho^{-3})\,.

For ρ0\rho\to 0, recall from the proof of Proposition 4 (i) that ρcoshx1log(1/ρ)\rho\cosh x_{1}\sim\log(1/\rho)\to\infty, which gives the asymptotics

(123) g~2(ρ)=161log(1/ρ)+O(log2(1/ρ)).\tilde{g}_{2}(\rho)=-\frac{1}{6}\frac{1}{\log(1/\rho)}+O(\log^{-2}(1/\rho))\,.

References

  • [1] A. Antonov, M. Konikov and M. Spector, Modern SABR Analytics. Springer, New York, 2019.
  • [2] L.P. Arguin, N.-L. Liu and T.-H. Wang. (2018) Most-likely-path in Asian option pricing under local volatility models. International Journal of Theoretical and Applied Finance. 21, 1850029.
  • [3] P. Barrieu, A. Rouault and M. Yor, A study of the Hartman-Watson distribution motivated by numerical problems related to the pricing of Asian options, J. Appl. Probab. 41, 1049-1058 (2004)
  • [4] G. Bernhart and J.-F. Mai, A note on the numerical evaluation of the Hartman-Watson density and distribution function, in Innovations in Quantitative Risk Management, p.337, K. Blau, M. Scherer and R. Zagst (Eds.), Springer Proceedings in Mathematics and Statistics, Springer, 2015.
  • [5] P. Boyle and A. Potapchik, Prices and sensitivities of Asian options: A survey. Insurance: Mathematics and Economics 42, 189-211 (2008)
  • [6] N. Cai, Y. Song and N. Chen, Exact simulation of the SABR model, Operations Research 65(4), 931-951 (2017)
  • [7] A. Comtet, C. Monthus and M. Yor, Exponential functionals of Brownian motion and disordered systems. J. Appl. Probab. 35, 255 (1998)
  • [8] D. Dufresne, Laguerre series for Asian and other options, Math. Finance 10, 407-428 (2000)
  • [9] D. Dufresne, The integral of the geometric Brownian motion, Adv. Appl. Probab. 33, 223-241 (2001).
  • [10] D. Dufresne, Bessel processes and a functional of Brownian motion, in M. Michele and H. Ben-Ameur (Eds.), Numerical Methods in Finance 35-57, Springer, 2005.
  • [11] A. Erdélyi, Asymptotic expansions, Dover Publications 1956
  • [12] S. Gerhold, The Hartman-Watson distribution revisited: asymptotics for pricing Asian options, J. Appl. Probab. 48(3), 892-899 (2011).
  • [13] A. Gulisashvili and E.M. Stein, Asymptotic behavior of the distribution of the stock price in models with stochastic volatility: the Hull-White model, C.R. Math. Acad. Sci. Paris 343, 519-523 (2006)
  • [14] A. Gulisashvili and E.M. Stein, Asymptotic behaviour of distribution densities in models with stochastic volatility. I. Math. Finance 20, 447-477 (2010)
  • [15] P. Hartman and G.S. Watson, ”Normal” distribution functions on spheres and the modified Bessel functions, Ann. Probab. 2, 593-607 (1974)
  • [16] K. Ishyiama, Methods for evaluating density functions of exponential functionals represented as integrals of geometric Brownian motion, Methodol. Comput. Appl. Probab. 7, 271-283 (2005)
  • [17] J.T. Kent, The spectral decomposition of a diffusion hitting time, Ann. Probab. 10(1), 207-219 (1982).
  • [18] H. Matsumoto and M. Yor, On Dufresne’s relation between the probability laws of exponential functionals of Brownian motions with different drifts, Adv. Appl. Probab. 35, 184-206 (2003)
  • [19] H. Matsumoto and M. Yor, Exponential functionals of Brownian motion: I. Probability laws at fixed time, Prob. Surveys 2, 312-347 (2005)
  • [20] F.W.J. Olver, Introduction to asymptotics and special functions, Academic Press 1974.
  • [21] D. Pirjol, L. Zhu, Short maturity Asian options in local volatility models, SIAM J. Finan. Math. 7(1), 947-992 (2016); arXiv:1609.07559[q-fin].
  • [22] D. Pirjol, L. Zhu, Asymptotics for the discrete-time average of the geometric Brownian motion, Adv. Appl. Probab. 49, 1-53 (2017)
  • [23] Pirjol, D. and L. Zhu. (2017). Short maturity Asian options for the CEV model, Probab. Eng. Inform. Sc. 33(2), 258-290, 2019. arXiv:1702.03382 [q-fin.PR]
  • [24] M. Schröder, On the integral of the geometric Brownian motion, Adv. Appl. Probab. 35, 159-183 (2003)
  • [25] M. Yor, On some exponential functionals of Brownian motion, Adv. Appl. Probab. 24, 509-531 (1992)