12

Why is

$$f(t) = \frac{1}{2πj}\int_{\sigma-j\infty}^{\sigma+j\infty} F(s) e^{st} \, ds,$$

provided that

$$F(s) = \int_{0}^{\infty} f(t) e^{-st} \, dt \ ?$$

I tried to find out myself, or searched online and found a term Bromwich integral, but I want to know how this expression is derived. (And I couldn't find any :()

Thank you.

Sangchul Lee
  • 181,930
  • You missed the factor $\frac{1}{2\pi i}$ in front of the Bromwich integral. – Sangchul Lee Sep 22 '18 at 17:50
  • @SangchulLee Oh, thank you, edited. – Hyeonseo Yang Sep 22 '18 at 23:22
  • For the History of this, which is substantial, I suggest reading Operational Methods in Applied Mathematics by H. S. Carslaw. It's an inexpensive Dover publication now. Here's a link at Amazon: https://www.amazon.com/Operational-methods-applied-mathematics-advanced/dp/B0006P59T6/ref=sr_1_4?ie=UTF8&qid=1538934907&sr=8-4&keywords=H.+S.+Carslaw – Disintegrating By Parts Oct 07 '18 at 17:56
  • @DisintegratingByParts Thank you! I will add that to my to-do list after CSAT. – Hyeonseo Yang Oct 07 '18 at 22:14

2 Answers2

24

It is the Fourier inversion formula in disguise. In case you have never encountered this theorem before, let me prove the following version (which is obviously far from optimal).

Proposition. Let $F(s) = \int_{0}^{\infty} f(t)e^{-st} \, dt$ be the Laplace transform of $f : [0,\infty) \to \mathbb{R}$. Assume that the following technical conditions hold with some $g : [0,\infty) \to \mathbb{R}$ and $\sigma \in \mathbb{R}$:

  • $f(t) = f(0) + \int_{0}^{t} g(u) \, du$. (In particular, $g$ is the 'derivative' of $f$.)
  • Both $f(t)e^{-\sigma t}$ and $g(t)e^{-\sigma t}$ are Lebesgue-integrable on $[0, \infty)$.

Then for any $s > 0$, we have $$ \lim_{R\to\infty} \frac{1}{2\pi i} \int_{\sigma-iR}^{\sigma+iR} F(z)e^{s z} \, dz = f(s). $$

Proof. Define $S(x) = \frac{1}{2} + \frac{1}{\pi}\int_{0}^{x} \frac{\sin t}{t} \, dt$. Then $S(x)$ is bounded, and by Dirichlet integral, we have

$$ \lim_{R\to\infty} S(Rx) = H(x) := \begin{cases} 1, & x > 0 \\ \frac{1}{2}, & x = 0 \\ 0, & x < 0 \end{cases} $$

(Obviously $H$ denotes the Heaviside step function.) Now we have

\begin{align*} \frac{1}{2\pi i} \int_{\sigma-iR}^{\sigma+iR} F(z)e^{s z} \, dz &= \frac{1}{2\pi} \int_{-R}^{R} F(\sigma + i\xi)e^{s(\sigma+i\xi)} \, d\xi \\ &= \frac{1}{2\pi} \int_{-R}^{R} \left( \int_{0}^{\infty} f(t)e^{-(\sigma+i\xi)t} \, dt \right)e^{s(\sigma+i\xi)} \, d\xi. \end{align*}

By Fubini's theorem, we can interchange the order of integral to obtain

\begin{align*} \frac{1}{2\pi i} \int_{\sigma-iR}^{\sigma+iR} F(z)e^{s z} \, dz &= \int_{0}^{\infty} f(t)e^{-(t-s)\sigma} \left( \frac{1}{2\pi} \int_{-R}^{R} e^{(s-t)i\xi} \, d\xi \right) \, dt \\ &= \int_{0}^{\infty} f(t)e^{-(t-s)\sigma} \left( \frac{\sin R(t-s)}{\pi (t-s)} \right) \, dt \end{align*}

By the assumption, both $f(t)e^{-\sigma t}$ and $(f(t)e^{-\sigma t})' = (f'(t) - \sigma f(t))e^{-\sigma t}$ are Lebesgue-integrable. In particular, this tells that $f(t)e^{-\sigma t}$ converges to $0$ as $t\to\infty$. So by integration by parts,

\begin{align*} \frac{1}{2\pi i} \int_{\sigma-iR}^{\sigma+iR} F(z)e^{s z} \, dz &= - f(0)e^{s\sigma} S(-Rs) - \int_{0}^{\infty} (f(t)e^{-(t-s)\sigma})' S(R(t-s)) \, dt. \end{align*}

As $R \to \infty$, the right-hand side converges to

\begin{align*} \lim_{R\to\infty} \frac{1}{2\pi i} \int_{\sigma-iR}^{\sigma+iR} F(z)e^{s z} \, dz &= - \int_{0}^{\infty} (f(t)e^{-(t-s)\sigma})' H(t-s) \, dt \\ &= - \left[ f(t)e^{-(t-s)\sigma} \right]_{t=s}^{t=\infty} = f(s). \end{align*}

(Pushing the limit inside the integral is justified by the dominated convergence theorem.)

Sangchul Lee
  • 181,930
  • 1
    Thank you for your elaborate answer! – Hyeonseo Yang Sep 23 '18 at 06:19
  • @sangchullee An integrable function $f(x)$ need not approach $0$ as $x\to \infty$. So, do we need an additional condition on $f$ here? – Mark Viola Mar 24 '19 at 19:41
  • 2
    @MarkViola, That's another place the integrability of $f'(x)e^{-\sigma x}$ enters. Indeed, notice that $$ f(b)e^{-\sigma b} - f(a)e^{-\sigma a} = \int_{a}^{b} \left( f(t) e^{-\sigma t} \right)' , \mathrm{d}t = \int_{a}^{b} (f'(t) - \sigma f(t)) e^{-\sigma t} , \mathrm{d}t $$ and that $ t \mapsto (f'(t) - \sigma f(t)) e^{-\sigma t}$ is integrable. So, as $a, b \to \infty$, this converges to zero, and by the Cauchy criterion, $f(x)e^{-\sigma x} $ converges as $x \to \infty$. The limiting value is automatically determined as $0$ by the integrability of $f(x)e^{-\sigma x} $. – Sangchul Lee Mar 24 '19 at 19:45
  • To simplify a bit, let's take $f(x)=e^{\sigma x}h(x)$. Then, by assumption $h$ is integrable. So, by the Cauchy Criterion, $\lim_{a,b\to\infty}\int_a^b h'(x),dx=\lim_{a,b\to\infty}(h(b)-h(a))=0$.

    While it is evident that $\lim_{x\to \infty}h(x)$ exists since $\lim_{b\to \infty}\int_a^b h'(x),dx=\lim_{b\to \infty}(h(b)-h(a))$ exists, how does one conclude that $\lim_{x\to \infty}h(x)=0$.

    – Mark Viola Mar 24 '19 at 21:13
  • @MarkViola, The trick is rather simple: if $\lim_{x\to\infty}h(x) $ is other than zero, then $h$ cannot be integrable on $(0, \infty)$. – Sangchul Lee Mar 24 '19 at 21:26
  • That is not true. For example $\int_0^\infty \cos(x^2),dx$ exists as an improper Riemann integral, but $\lim_{x\to\infty}\cos(x^2)$ fails to exist. HERE are several other examples. – Mark Viola Mar 24 '19 at 22:24
  • @MarkViola, Ah, now I see why you object that part. I failed to clearly address that what I really intended is absolute-integrability (i.e. Lebesgue-integrability in this case) as is usual in analysis literature. I definitely agree that improper integrability is not sufficient for the conclusion. (In such case, we have to pay arbitrarily small amount of abscissa to retrieve absolute integrability.) I will update my answer to correctly address this. – Sangchul Lee Mar 24 '19 at 22:31
  • I thought that you might be thinking of Lebesgue integrability, but you wrote a principal value integral for the inverse Laplace Transform. That convinced me that we were talking about Riemann integrals. – Mark Viola Mar 24 '19 at 22:34
  • But that said, there are examples of functions that don't approach $0$ at infinity and are absolutely integrable as improper Riemann integrals. – Mark Viola Mar 24 '19 at 22:41
  • @MarkViola Thank you for pointing that out, I now updated my answer to clearly address which mode of convergence is involved. – Sangchul Lee Mar 24 '19 at 22:41
2

Here I present another version of the inversion formula for the Laplace Transform and a proof based entirely on the Fourier transform. This version extends the version described by Sangchul Lee.

Throughout this posting

  • $m$ denotes the Lebesgue measure on the real line.
  • For any function $f\in L^{loc}_1((0,\infty)$, its Laplace transform is defined as $$\bar{f}(s)=\int^\infty_0e^{-st}f(t)\,dt$$ If the integral above converges absolutely for some $s>0$, then $\bar{f}$ can be extended as an analytic function of the half place $H_{s}=\{z\in\mathbb{C}:\mathfrak{R}(z)>s\}$ that is continuous along the vertical line $\mathfrak{R}(z)=s$.
  • For any $g\in L_1(\mathbb{R})$, its Fourier transform is defined as $$\widehat{g}(\xi)=\int_\mathbb{R} e^{-2\pi I\xi x}g(x)\,dx$$ For any complex Borel measure (or real valued measure of total finite variation) $\mu$ on $\mathbb{R}$, its Fourier transform (or its characteristic function) is defined as $$\widehat{\mu}(\xi)=\int_\mathbb{R}e^{ix\xi}\mu(dx)$$ Notice that if $\mu\ll m$ and $\mu=g\cdot m$, then $\widehat{g}(\xi)=\widehat{\mu}(-2\pi \xi)$.

Inversion formulas for the Fourier transform are well known and have been discussed here at MSE. I will use the following version of the Fourier inversion formula.

Theorem ($L_1$ summability) Suppose $f\in L_1(m)$ and that $f$ is of bounded variation in a neighborhood of some point $x\in\mathbb{R}$. Then \begin{align} \frac{f(x-)+f(x+)}{2}=\frac{1}{2\pi}\lim_{T\rightarrow\infty}\int^T_{-T}e^{-itx} \widehat{f}(-t/2\pi)\,dt\tag{4}\label{four} \end{align} where $f(x-)=\lim_{y\nearrow x}f(y)$ and $f(x+)=\lim_{y\searrow x}f(y)$.

I provide a proof of this result at the end of my posting.

For example, if $f\in L_1(m)$ is piecewise continuously differentiable, then $f$ is local bounded variation and thus, \eqref{four} holds.


Derivation of inversion formula for Laplace Transform: Extend $f$ to $\mathbb{R}$ by setting $f(t)=0$ for $t\leq 0$. Suppose $g_c(t)=e^{-ct}f(t)\in L_1((0,\infty),m)$ for some $c>0$. It follows that $$\widehat{g_c}(\xi)=\int^\infty_{0}e^{-(c+2\pi\xi i)t}f(t)\,dt=\overline{f}(c+2\pi \xi i)$$ Then, by the summability theorem above, $$ \frac{g_c(t-)+g_c(t_+)}{2}=\frac1{2\pi}\lim_{T\rightarrow\infty}\int^T_{-T}e^{-it\xi}\widehat{g_c}(-\xi/2\pi)\,d\xi=\frac1{2\pi}\lim_{T\rightarrow\infty}\int^T_{-T}e^{it\xi}\overline{f}(c+i\xi)\,d\xi$$ at any point $t$ around which $g_c$ (equivalently $f$) is of bounded variation. Hence, for such point $t$, $$\frac{f(t-)+f(t-)}{2}=\frac1{2\pi}\lim_{T\rightarrow\infty}\int^T_{-T}e^{t(c+i\xi)}\overline{f}(c+i\xi)\,d\xi= \frac{1}{2\pi i}\lim_{T\rightarrow\infty}\int^{c+iT}_{c-iT}e^{tz}\overline{f}(z)\,dz $$


Proof of $L_1$ summability theorem: By Fubini's theorem we have that $$g(x):=\int^T_{-T}\widehat{f}(-t/2\pi) e^{-ixt}\,dt=\int_{\mathbb{R}}\int^T_{-T}f(y)e^{i(y-x)t}\,dt\,dy= \int_{\mathbb{R}}f(y)\frac{\sin(T(y-x))}{y-x}dy $$ Since $t\mapsto\frac{\sin t}{t}$ is even, it follows that $$g(x)=\int_{\mathbb{R}}f(y+x)\frac{\sin(Ty)}{y}dy=\int_{\mathbb{R}}f(x-y)\frac{\sin(Ty)}{y}dy$$

Suppose $f$ is of bounded variation in the interval $I_\delta=(x-\delta,x+\delta)$. With our loss of generality, we may assume that $f$ is monotone nondecreasing on $I_\delta$. Splitting the domain of integration gives \begin{align} g(x)&=\frac1\pi\int_{\mathbb{R}}\frac{f(y+x)+f(x-y)}{2}\frac{\sin Ty}{y}\,dy=\frac2\pi\int^\infty_0\frac{f(y+x)+f(x-y)}{2}\frac{\sin Ty}{y}\,dy\\ &=\frac2\pi\int^\delta_0\left(\frac{f(y+x)+f(x-y)}{2}-\frac{f(x-)+f(x+)}{2}\right)\frac{\sin Ty}{t}\,dy\\ &\quad + \frac2\pi\int^\infty_\delta\frac{f(y+x)+f(x-y)}{2}\frac{\sin Ty}{y}\,dy\\ &\quad + \frac{f(x-)+f(x+)}{2}\frac{2}{\pi}\int^\delta_0\frac{\sin Ty}{y}\,dy \end{align}

The third integral in the right converges to $\frac{f(x-)+f(x+)}{2}\frac{2}{\pi}\int^\infty_0\frac{\sin u}{u}\,du=\frac{f(x-)+f(x+)}{2}$.

The second integral in the right converges to $0$ by The Riemann-Lebesgue lemma, for $y\mapsto\frac{f(y+x)-f(x-y)}{y}\mathbb{1}_{(\delta,\infty)}(y)\in L_1(m)$.

The first integral on the right requires a little extra effort. Let $A=\sup_{y>0}\Big|\int^y_0\sin\frac{\sin y}{y}\,dy\Big|$. Define \begin{align} h(y;x):=\frac{f(y+x)+f(x-y)}{2}-\frac{f(x-)+f(x+)}{2}= \frac{f(y+x)-f(x+)}{2} +\frac{f(x-y)-f(x-)}{2} \end{align}

Given $\varepsilon>0$, there is $0<\eta<\delta$ such that $f(y)-f(x_+)<\frac{\varepsilon}{2A}$ for $x<y<x+\delta$, and $f(x-)-f(y)<\frac{\varepsilon}{2A}$ for $x-\delta<y<x$.
By the second mean value theorem for integrals there are points $\xi_+, \xi_-\in(0,\eta)$ such that \begin{align} \int^\eta_0h(y;x)\frac{\sin Ty}{y}\,dy&=\frac{f(\xi_+ x)-f(x-)}{2}\int^\eta_{\xi_+}\frac{\sin Ty}{y}\,dy+\frac{f(x-\xi_-)-f(x-)}{2}\int^\eta_{\xi_-}\frac{\sin Ty}{y}\,dy \end{align} It follows that $$\Big|\int^\eta_0h(y;x)\frac{\sin Ty}{y}\,dy\Big|<\varepsilon$$ On the other hand, $y\mapsto h(y;x)\mathbb{1}_{(\eta,\delta)}(y)\in L_1(m)$ and so, $\int^\delta_\eta h(y;x)\frac{\sin Ty}{y}\,dy$ converges to $0$ by the Riemann-Lebesgue lemma.

Putting things together we have that $$\limsup_{T\rightarrow\infty}\left|\int^T_{-T}\widehat{f}(-t/2\pi) e^{-ixt}\,dy - \frac{f(x-)+f(x-)}{2}\right|\leq\varepsilon$$ As $\varepsilon>0$ is arbitrary, the conclusion of the Theorem follows.

Mittens
  • 46,352