The Method of Steepest Descent and Asymptotic Forms of Airy Function
If you have read some textbooks or have taken any lectures pertaining
to the WKB approximation while studying quantum physics in undergraduate level,
then you might, with high probability, be irritated from the situation that both
textbooks and professors just apprise us the formulae of asymptotic forms of Airy
function without any sophisticated explanations. (Of course I know it is inevitable
bcz of deficient space of textbooks and insufficient time for explaning every
underlying mathematics in lectures. Hahaha...). Nevertheless, if you are a student
who majors in theoretical physics, then you are obliged to ask a question "WHY?"
albeit nobody tells you any underlying reasons!!!
Hence, in this post, we are going to derive an approximation method
which is called The Method of Steepest Descent (or Saddle-Point
Approximation), and will apply it to Airy function in purpose to
obtain its asymptotic forms.
1. The Method of Steepest Descent (Saddle-Point Approximation)
" How do we obtain an approximation form of a real-valued function $f(t)$ defined on a real domain? "
Let us write the real-valued function $f(t)$ in a form of a contour integral
along the contour $C$ \[f(t) = \int_{C} dz\, F(z,t)\] where $F$ is an analytic function and $\forall t \in \mathbb{R}$.
Since $F$ is analytic and an
exponential function preserves holomorphism, it can be expressed in
terms of an exponential function whose exponent is an analytic function.
i.e. \[F(z,t) = e^{w(z,t)}\] where
\[w(x,y) = u(x,y) + i \, v(x,y) \quad
\forall x,y \in \mathbb{R} \;\;.\]
Lemma. Both $u$ and $v$ cannot have extremum if $w = u + i\,v$ is
analytic.
proof. Assume that $u(x,y)$
has a local minimum(maximum) at $(x,y) =
(x_0, y_0)$. Then the Hessian matrix of $u(x,y)$ at $(x_0, y_0)$ should be
positively(negatively) definite, and this implies $\left.\frac{\partial^2 u}{\partial
x^2}\right|_{x_0,y_0}$ and $\left.\frac{\partial^2 u}{\partial
y^2}\right|_{x_0,y_0}$ have same signs. (If not, then the
determinant of Hessian matrix becomes negative and this violates the
condition of positively(negatively) definite Hessian matrix since
determinant should be positive in both cases.) Since $w(x,y)$ is analytic, however, $u(x,y)$ must satisfy ${\nabla}^2 u(x,y) = 0$. This implies $\left.\frac{\partial^2 u}{\partial
x^2}\right|_{x_0,y_0}$ and $\left.\frac{\partial^2 u}{\partial
y^2}\right|_{x_0,y_0}$ cannot have same signs, and we get a
contradiction. By applying the identical procedure to $v(x,y)$, we may show that there is NO
extremum for $u$ and $v$ if $w = u +
i v$ is analytic.$\quad
\blacksquare$
Although $u(x,y)$ and $v(x,y)$ cannot have extremum, both can
have a saddle point(say $z_0$), and
this point has a crucial rule for developing this methodology.
Let us expand $w(z,t)$ near the saddle point $z=z_0$ up to second order of $z$. \[w(z,t) = w(z_0,t) + \frac{1}{2}
\left.\frac{\partial^2 w}{\partial z^2}\right|_{z_0} (z-z_0)^2 +
\cdots \] Note that $\left.\frac{dw}{dz}\right|_{z_0}$ is vanished since $z_0$ is a saddle point.
For convenience, let us introduce some abbreviations, and let’s denote
the phases explicitly. \[w(z_0,t) = w_0 , \quad \left.\frac{\partial^2
w}{\partial z^2}\right|_{z_0} = w''_0 =
\left|{w''_0}\right| e^{i\alpha} , \quad z-z_0 = re^{i\theta}
\;\;.\]
Then we can rewrite the $w(z,t)$ as
\[\begin{aligned}
w(z,t) &= w_0 + \frac{1}{2}\left|w''_0\right|
e^{i\alpha} (r e^{i\theta})^2 + \mathrm{O}(r^3) \\ &= w_0 +
\frac{1}{2}\left|w''_0\right| r^2 e^{ i(\alpha + 2\theta) } +
\mathrm{O}(r^3) \\ &= w_0 + \frac{1}{2} \left|w''_0\right|
r^2 \left[ \cos(\alpha + 2\theta) + i \, \sin(\alpha + 2\theta) \right]
+ \mathrm{O}(r^3) \\ &\approx \left[ w_0 + \frac{1}{2}
\left|w''_0\right| r^2 \cos(\alpha + 2\theta) \right] + i \,
\left[ \frac{1}{2} \left|w''_0\right| r^2 \sin(\alpha + 2\theta)
\right] \;\;.
\end{aligned}\]
Remember that we have defined $w(z,t)$ by $w =
u + i v$. Hence we obtain \[\begin{cases}
\quad u(z,t) = w_0 + \frac{1}{2} \left|w''_0\right| r^2
\cos(\alpha + 2\theta) \\ {} \\ \quad v(z,t) = \frac{1}{2}
\left|w''_0\right| r^2 \sin(\alpha + 2\theta)
\end{cases}\]
Consider the situation in which $z$
departs from $z_0$ (i.e. as r
increases from 0).
$\mathbf{\cdot}$ For $u(z,t)$
$\qquad$ i) $u$ will increase most rapidly when \[\cos(\alpha + 2\theta) = 1 \quad \Rightarrow
\quad \alpha + 2\theta = 2n\pi \quad \Rightarrow \quad \theta =
-\frac{\alpha}{2} +n\pi\] $\qquad$ ii) $u$ will decrease most rapidly when \[\cos(\alpha + 2\theta) = -1 \quad \Rightarrow
\quad \alpha + 2\theta = (2n+1)\pi \quad \Rightarrow \quad \theta =
-\frac{\alpha}{2} + \left(n+\frac{1}{2}\right)\pi\] $\qquad$ iii) $u$ will be invariant when \[\cos(\alpha + 2\theta) = 0 \quad \Rightarrow
\quad \alpha + 2\theta = \left(n+\frac{1}{2}\right)\pi \quad \Rightarrow
\quad \theta = -\frac{\alpha}{2} +
\left(\frac{n}{2}+\frac{1}{4}\right)\pi\]
$\mathbf{\cdot}$ For $v(z,t)$
$\qquad$ i) $v$ will increase most rapidly when \[\sin(\alpha + 2\theta) = 1 \quad \Rightarrow
\quad \alpha + 2\theta = \left(2n+\frac{1}{2}\right)\pi \quad
\Rightarrow \quad \theta = -\frac{\alpha}{2}
+\left(n+\frac{1}{4}\right)\pi\] $\qquad$ ii) $v$ will decrease most rapidly when \[\sin(\alpha + 2\theta) = -1 \quad \Rightarrow
\quad \alpha + 2\theta = \left(2n+\frac{3}{2}\right)\pi \quad
\Rightarrow \quad \theta = -\frac{\alpha}{2} +
\left(n+\frac{3}{4}\right)\pi\] $\qquad$ iii) $v$ will be invariant when \[\sin(\alpha + 2\theta) = 0 \quad \Rightarrow
\quad \alpha + 2\theta = n\pi \quad \Rightarrow \quad \theta =
-\frac{\alpha}{2} + \frac{n\pi}{2}\]
Let us discuss about some crucial notions on performing the integral.
How do we choose an optimum integration contour?
We need to choose a contour that passes through the saddle point $z_0$ in the directions of maximum rate of decrease in $u$ with distance from $z_0$, and therefore also in $\vert F \vert$ (since $\vert F \vert = e^{u}$). These directions are level lines of $v(x,y)$, so the phase factor of $F$ (i.e. $e^{iv}$) will not change as we leave tha saddle point. If we chose $z_0$ to be a point other than a saddle point, then the expansion of $w$ (eqn. (1)) would have contained a non-zero linear term in $r$ (since $\left.\frac{\partial u}{\partial z}\right|_{z_0}$ does not vanish.), and it might be impossible to construct a curve which passes through $z_0$ that makes $\vert F \vert = e^u$ to decrease and keeps the phase of F (i.e. $e^{iv}$) constant, as $z$ departs from $z_0$.
Evaluation of the integral
Let us assume that ${\vert w''_0 \vert}$, the measure of the rate of decrease in $\vert F \vert$ as we leave $z_0$, is large enough that the bulk of the value of the integral has already been attained for small $R$. (i.e. The significant contribution to the integral only comes from the immediate vicinity of the saddle point $z_0$.)
Split the integral as below. \[\begin{aligned} f(t) \;=\; \int_{C} dz\,F(z,t) \; &= \; \int_{\text{near the saddle point}} dz \, F(z,t) \;+\; \int_{\text{error contribution}} dz \, F(z,t) \\ &{} \\ &= \; \int_{\text{descend from } F(z_0)} dz \, F(z,t) \;+\; \int_{\text{ascend to } F(z_0)} dz \, F(z,t) \;+\; \int_{\text{error contribution}} dz \, F(z,t) \end{aligned}\]
Near the saddle point, $z = z_0 + r e^{i\theta}$ gives $dz = e^{i\theta} dr$ since we assumed
that "the significant contribution to the integral only comes from the immediate vicinity of the saddle point" and
this allows us to approximate the integration contour be a straight line right near the saddle point without making significant error. Hence,
\[f(t) \;=\; \int_{0}^{R} dr \, r^2
e^{i\theta} \exp{ \left( w_0 + \frac{1}{2} {\vert w''_0 \vert}
e^{i(\alpha+2\theta)} \right)} \;+\; \int_{R}^{0} dr \, r^2
e^{i\bar{\theta}} \exp{ \left( w_0 + \frac{1}{2} {\vert w''_0
\vert} e^{i(\alpha+2\bar{\theta})} \right)} \;+\; \text{error
term}\] where $\bar{\theta} = \theta -
\pi$.
Since we’ve set the integration contour to satisfy $e^{i(\alpha + 2\theta)} = e^{i(2n+1)\pi} =
-1$ right near the saddle point, \[\begin{aligned}
f(t) \;&=\; e^{w_0 + i\theta} \int_{0}^{R} dr \, r^2 e^{ -
\frac{1}{2} {\vert w''_0 \vert}^2 } \;+\; e^{w_0 +
i(\theta-\pi)} \int_{R}^{0} dr \, r^2 e^{ - \frac{1}{2} {\vert
w''_0 \vert} } \;+\; \text{error term} \\ &=\; 2 \, e^{w_0 +
i\theta} \int_{0}^{R} dr\, e^{-\frac{1}{2} {\vert w''_0 \vert}
r^2 } \;+\; \text{error term} \;\;.
\end{aligned}\]
Note that our assumption enables us to replace $R \to \infty$ without making any significant error.
\[f(t) \;\approx \; 2\, e^{w_0 + i\theta}
\int_{0}^{\infty} dr \, e^{-\frac{1}{2} {\vert w''_0 \vert} r^2}
\;=\; 2\,e^{w_0 + i\theta} \frac{1}{2} \sqrt{ \frac{2\pi}{\vert
w''_0 \vert} } \;=\; e^{w(z_0,t)} e^{i\theta} \sqrt{
\frac{2\pi}{\vert{ w''(z_0,t) \vert} } }\]
Therefore, for $\vert t \vert \to
\infty$, we can approximate the $f(t)$ by \[\therefore \qquad f(t) \overset{|t| \to
\infty}{\approx} F(z_0,t) \, e^{i\theta} \sqrt{ \frac{2\pi}{\vert
w''(z_0,t) \vert} }\] where $z_0$ is a saddle point of $w(z,t)$ and \[\theta = -\frac{1}{2}
\mathrm{arg}{\left[w''(z_0,t)\right]} + { \left( { \frac{\pi}{2}
\text{ or } \frac{3\pi}{2} } \right)\,. }\] Choice of $\theta$ (which affects only the sign of
the final result) is determined from the sense in which the contour
passes through the saddle point $z_0$.
Let’s derive asymptotic forms of Airy function applying this
approximation method!
2. The Asymptotic Form of Airy Function
To apply the saddle point approximation, first we need to express Airy function in a form of complex contour integral.
Consider a differential equation whose solutions are Airy functions.
\[\begin{aligned}
\frac{d^2 f}{d t^2} = t\, f(t) \tag{1}
\end{aligned}\]
Let us write the solution $f(t)$
in a form of \[f(t) = \int_{C} dz \,
h(z)\,e^{tz} \quad \forall z \in \mathbb{C}
\;. \tag{2}\]
By plugging (2) into (1), L.H.S. becomes \[\frac{d^2 f}{d t^2} = \int_{C} dz \,
h(z)\frac{d^2}{d t^2}e^{tz} = \int_{C} dz\, z^2 h(z) e^{tz} \;,\]
and by integrating by parts, R.H.S. becomes \[t\,f(t) = \int_{C} dz\, h(z)\, t\, e^{tz} =
\left[\, h(z)\,e^{tz} \,\right]_{\text{ends of } C} - \int_{C}
dz\,e^{tz}\frac{dh}{dz} \;.\]
In order to make sure the equality in eqn(1), we must set \[\left[\, h(z)\,e^{tz} \,\right]_{\text{ends of }
C} = 0 \;.\]
Accepting the constraint above, we can express $h(z)$ as \[z^2
h(z) = -\frac{dh}{dz} \quad \Rightarrow \quad h(z) = c \cdot
e^{-\frac{1}{3}z^3} \; .\]
Since we’ve set the square bracket term to be annihilated, each end
point of the integration contour $C$
must satisfy \[\exp{\left[ -\frac{1}{3}
{z^3_\text{end}} + t\, z_{\text{end}} \right]} = 0 \;.\]
This condition implies \[\mathrm{Re}{(z^3_{\text{end}})} > 0 \qquad
\text{and} \qquad \left| \, z_{\text{end}} \, \right| \to
\infty\] because $z^3$
diverges much faster than $z$, as
$|z|$ increases.
From $\mathrm{Re}{(z^3_{\text{end}})}>0$, we
obtain \[\frac{2n\pi}{3} - \frac{\pi}{6} \,
< \, \arg{(z_{\text{end}})} \,<\, \frac{2n\pi}{3} + \frac{\pi}{6}
\tag{3}\]
and according to this constraint, the complex plane is divided as below.
Due to the condition (3), we must set our integration curve $C$ to start from an infinity in one gray
sector, and terminate at infinity in the other gray sector. Note also
that since there are three different sectors that satisfy the condition
(3), we can make three different contours, say $C_1, C_2, C_3$, and each choice will give
corresponding solutions $f_1, f_2,
f_3$.
Consequently, the general solutions(in the contour integral form) to the differential equation (1) can be written as \[\begin{aligned}
& \mathrm{Ai}\,(t) = \frac{1}{2\pi i} \int_{C_1}
dz\,e^{zt-\frac{z^3}{3}} \\ &{} \\ &
\mathrm{Bi}\,(t) = \frac{1}{2\pi} \int_{C_2 - C_3} dz\,
e^{zt-\frac{z^3}{3}}
\end{aligned}\]
2.1 Asymptotic form of $\mathrm{Ai}\,(t)$ in $\mathbf{t \gg 0}$ region
First, we should figure out saddle points of $w$. \[e^{w(z,t)} = e^{tz-\frac{z^3}{3}} \quad
\Rightarrow \quad \frac{dw}{dz}=t-z^2 \quad \Rightarrow \quad
\text{saddle points of } w(z,t) \text{ : } \;\; z_0 = \pm
\sqrt{t}\]
For each saddle point, let’s inspect in which direction $\mathrm{Re}(w)$ decreases most
rapidly.
For $z_0 = -\sqrt{t}$
$w''(z_0,t) = 2\sqrt{t}$
$\Rightarrow \quad \mathrm{arg}(w''_0) = 0$
$\Rightarrow \quad \mathrm{Re}(w)$ decreases most rapidly when $\theta = \pi / 2 \, , \; 3\pi / 2$ as $z$ leaves $z_0$.
For $z_0 = \sqrt{t}$
$w''(z_0,t) = -2\sqrt{t}$
$\Rightarrow \quad \mathrm{arg}(w''_0) = \pi$
$\Rightarrow \quad \mathrm{Re}(w)$ decreases most rapidly when $\theta = 0 \, , \; \pi$ as $z$ leaves $z_0$.
Let us definitize the integration contour in an explicit algebraic
form. To do so, All we need to do is to find the level lines of $v$ that pass through each saddle point of
$w$.
By plugging $z=x+iy$ into $(w(z)$, we obtain \[w(z) = tz-\frac{1}{3}z^3 = \left[ \,
tx-\frac{1}{3}x^3 - xy^2 \, \right] + i\,\left[ \, ty + \frac{1}{3}y^3 -
x^2 y \, \right] \;.\]
Since $v(x,y)$ should be a level
line, \[v(x,y) = ty + \frac{1}{3}y^3 - x^2 y
= \gamma \qquad \text{where} \; \gamma \; \text{is a real
constant}\] and it must pass through the saddle point(s), \[v(z_0) = v(\pm \sqrt{t}, 0) = 0 = \gamma \;
.\]
Then, we obtain some candidates for the integration contour $C$. \[ty + \frac{1}{3}y^3 - x^2 y = 0 \qquad \implies \qquad C \; : \; \begin{cases} \quad y=0 & \mbox{: real axis} \\ {} \\ \quad x^2 - \frac{1}{3}y^2 = t & \mbox{: hyperbola} \end{cases}\]
However, the real axis cannot be an integration contour since its
left-end is not located inside the gray color sectors! Furthermore, the
hyperbola lies in $x>0$ region
should be neglected as well since its both end-points are not inside the
gray sectors.
Hence, the integration contour should be chosen as below. \[C = \left\{ z=x+iy \;\; \left| \quad
x<0 \; , \; y \in \mathbb{R} \quad \text{s.t.} \quad x^2 -
\frac{1}{3}y^2 = t \; \right. \right\}\]
Therefore, the integration contour for this case might be drawn as
below.
I chose the direction of the contour be upward in order to ensure that the derived asymptotic form of Airy function matches with the conventional definition of Airy function.
Plugging \[{ z_0 = -\sqrt{t} \; , \;\; w_0 = -\frac{2}{3}t^{\frac{3}{2}} \; , \;\; w''_0 = 2\sqrt{t} \; , \;\; \theta = -\frac{1}{2}\mathrm{arg}(w''_0) + \frac{\pi}{2} = \frac{\pi}{2} }\] into the general formula of the method of steepest descent \[\mathrm{Ai}\,(t) \sim \frac{1}{2\pi i} F(z_0,t) \, e^{i\theta} \sqrt{\frac{2\pi}{|w''(z_0,t)|}} \; ,\] we finally obtain the asymptotic approximation of $\mathrm{Ai}\,(t)$ in $t \gg 0$ region,
\[\therefore \qquad \mathrm{Ai}\,(t) \,
\overset{t \to \infty}{\sim} \, \frac{1}{2\sqrt{\pi}\, t^{\frac{1}{4}}}
\, \exp{\left(-\frac{2}{3} t^{\frac{3}{2}}\right)} \; .\]
2.2 Asymptotic form of $\mathrm{Ai}\,(t)$ in $\mathbf{t \ll 0}$ region
For this case, since $t$ is negative, saddle points are given by \[z_0 = \pm i \sqrt{|t|} \;.\]
For each saddle point, let’s inspect in which direction $\mathrm{Re}(w)$ decreases most rapidly.
For $z_0 = -i\sqrt{|t|}$
$w''(z_0,t) = 2i\sqrt{|t|}$
$\Rightarrow \quad \mathrm{arg}(w''_0) = \frac{\pi}{2}$
$\Rightarrow \quad \mathrm{Re}(w)$ decreases most rapidly when $\theta = \pi / 4 \, , \; 5\pi / 4$ as $z$ leaves $z_0$.
For $z_0 = i\sqrt{|t|}$
$w''(z_0,t) = -2i\sqrt{|t|}$
$\Rightarrow \quad \mathrm{arg}(w''_0) = -\frac{\pi}{2}$
$\Rightarrow \quad \mathrm{Re}(w)$ decreases most rapidly when $\theta = 3\pi / 4 \, , \; 7\pi / 4$ as $z$ leaves $z_0$.
For $z_0 = i\sqrt{|t|}$, the level line of $v$ which passes through the saddle this point is \[v(x,y) = ty + \frac{1}{3}y^3 - x^2 y = v(0, \sqrt{|t|}) \qquad \Rightarrow \qquad -|t|\,y + \frac{1}{3}y^3 - x^2 y = -\frac{2}{3} |t|^{\frac{3}{2}} \;\; .\]
Similarly, for $z_0 =
-i\sqrt{|t|}$, the level line of $v$ that passes through this point is \[-|t|\,y + \frac{1}{3}y^3 - x^2 y = \frac{2}{3}
|t|^{\frac{3}{2}} \;\; .\]
By picking appropriate level lines (which end-points are inside the gray
sectors), and by orienting the picked level lines properly, the
integration contours can be depicted as below.
Following the process analogous to the antecedent case, we obtain the asymptotic form of Airy function in $t \ll 0$ region.
\[\mathrm{Ai}\,(t) \;\; \sim \;\;
\frac{1}{2\pi i} \; \left[ \, e^{\frac{2\,i}{3} |t|^{\frac{3}{2}}}
e^{\frac{\pi}{4} i} \sqrt{\frac{2\pi}{\left|2i \sqrt{|t|}\right|}} \; +
\; e^{-\frac{2\,i}{3} |t|^{\frac{3}{2}}} e^{\frac{3\pi}{4} i}
\sqrt{\frac{2\pi}{\left|-2i \sqrt{|t|}\right|}} \, \right] \;\; = \;\;
\frac{1}{\sqrt{\pi} \, |t|^{\frac{1}{4}}} \; \sin{\left( \frac{2}{3}
|t|^{\frac{3}{2}} + \frac{\pi}{4} \right)}\]
\[\therefore \qquad \mathrm{Ai}\,(t) \;\;
\overset{t \to -\infty}{\sim} \;\; \frac{1}{\sqrt{\pi} \,
|t|^{\frac{1}{4}}} \;\sin{\left( \frac{2}{3} |t|^{\frac{3}{2}} +
\frac{\pi}{4} \right)}\]
Note that, if we included only one saddle point or if we set the
directions of contours to be opposite to each other, then $\mathrm{Ai}\,(t)$ becomes an
imaginary-valued function. Hence, in order to ensure that $\mathrm{Ai}\,(t)$ be real, then we should
take into account every saddle point, and should set the contours’
orientations as fig.4.
3. Conclusion
From the adhered figure above, we can observe that the asymptotic
forms well mimic the original Airy function without significant error
barring the range $|t|≪1$ in
which the asymptotic forms diverge.
References
James Brown. Complex Variables and Application. McGraw Hill
George B. Arfken. Mathematical Methods for Physicists. Academic Press
K. F. Riley. Mathematical Methods for Physics and Engineering. Cambridge University Press