Skip to main content

Fourier Analysis, Distributions, and Constant-Coefficient Linear PDE

  • Chapter
  • First Online:
  • 11k Accesses

Part of the book series: Applied Mathematical Sciences ((AMS,volume 115))

Abstract

Fourier analysis is perhaps the most important single tool in the study of linear partial differential equations. It serves in several ways, the most basic–and historically the first–being to give specific formulas for solutions to various linear PDE with constant coefficients, particularly the three classics, the Laplace, wave, and heat equations:

$$\Delta u = f,\quad\frac{{\partial }^{2}u} {\partial {t}^{2}} -\Delta u = f,\quad\frac{\partial u} {\partial t} -\Delta u = f,$$
(0.1)

with Δ = 2∂x12 + ⋯ + 2∂xn2. The Fourier transform accomplishes this by transforming the operation of ∂xj to the algebraic operation of multiplication by iξj. Thus the (0.1) are transformed to algebraic equations and to ODE with parameters.

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

Buying options

Chapter
USD   29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD   84.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD   109.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Learn about institutional subscriptions

References

  1. L. Bers, F. John, and M. Schechter, Partial Differential Equations, Wiley, New York, 1964.

    MATH  Google Scholar 

  2. R. Churchill, Fourier Series and Boundary Value Problems, McGraw Hill, New York, 1963.

    MATH  Google Scholar 

  3. J. Dadok and M. Taylor, Local solvability of constant coefficient linear PDE as a small divisor problem, Proc. AMS 82(1981), 58–60.

    Google Scholar 

  4. W. Donoghue, Distributions and Fourier Transforms, Academic, New York, 1969.

    MATH  Google Scholar 

  5. A. Erdelyi, Asymptotic Expansions, Dover, New York, 1965.

    MATH  Google Scholar 

  6. A. Erdelyi et al., Higher Transcendental Functions, Vols. I–III (Bateman Manuscript Project), McGraw-Hill, New York, 1953.

    Google Scholar 

  7. G. Folland, Introduction to Partial Differential Equations, Princeton University Press, Princeton, N. J., 1976.

    MATH  Google Scholar 

  8. F. G. Friedlander, Sound Pulses, Cambridge University Press, Cambridge, 1958.

    MATH  Google Scholar 

  9. A. Friedman, Generalized Functions and Partial Differential Equations, Prentice-Hall, Englewood Cliffs, N. J., 1963.

    Book  Google Scholar 

  10. I. M. Gelfand and I. Shilov, Generalized Functions, Vols. 1–3, Fizmatgiz, Moscow, 1958.

    Google Scholar 

  11. R. Goldberg, Fourier Transforms, Cambridge University Press, Cambridge, 1962.

    MATH  Google Scholar 

  12. E. Hille and R. Phillips, Functional Analysis and Semi-groups, Colloq. Publ. AMS, Providence, R.I., 1957.

    MATH  Google Scholar 

  13. E. Hobson, The Theory of Spherical Harmonics, Chelsea, New York, 1965.

    Google Scholar 

  14. L. Hörmander, The Analysis of Linear Partial Differential Equations, Vols. 1–2, Springer, New York, 1983.

    Google Scholar 

  15. L. K. Hua, Introduction to Number Theory, Springer, New York, 1982.

    Google Scholar 

  16. F. John, Partial Differential Equations, Springer, New York, 1975.

    Book  Google Scholar 

  17. Y. Katznelson, An Introduction to Harmonic Analysis, Dover, New York, 1976.

    MATH  Google Scholar 

  18. O. Kellogg, Foundations of Potential Theory, Dover, New York, 1954.

    MATH  Google Scholar 

  19. T. Körner, Fourier Analysis, Cambridge University Press, Cambridge, 1988.

    Book  Google Scholar 

  20. E. Landau, Elementary Number Theory, Chelsea, New York, 1958.

    MATH  Google Scholar 

  21. N. Lebedev, Special Functions and Their Applications, Dover, New York, 1972.

    MATH  Google Scholar 

  22. J. Leray, Hyperbolic Differential Equations, Princeton University Press, Princeton, N. J., 1952.

    Google Scholar 

  23. S. Mizohata, The Theory of Partial Differential Equations, Cambridge University Press, Cambridge, 1973.

    MATH  Google Scholar 

  24. F. Olver, Asymptotics and Special Functions, Academic, New York, 1974.

    MATH  Google Scholar 

  25. S. Patterson, An Introduction to the Theory of the Riemann Zeta-Function, Cambridge University Press, Cambridge, 1988.

    Book  Google Scholar 

  26. M. Pinsky and M. Taylor, Pointwise Fourier inversion: a wave equation approach, J. Fourier Anal. 3 (1997), 647–703.

    Article  MathSciNet  Google Scholar 

  27. W. Press, B. Flannery, S. Teukolsky, and W. Vetterling, Numerical Recipes; the Art of Scientific Computing, Cambridge University Press, Cambridge, 1986.

    MATH  Google Scholar 

  28. H. Rademacher, Lectures on Elementary Number Theory, Chelsea, New York, 1958.

    MATH  Google Scholar 

  29. M. Reed and B. Simon, Methods of Mathematical Physics, Academic, New York, Vols. 1,2, 1975; Vols. 3,4, 1978.

    Google Scholar 

  30. F. Riesz and B. S. Nagy, Functional Analysis, Ungar, New York, 1955.

    MATH  Google Scholar 

  31. W. Rudin, Real and Complex Analysis, McGraw-Hill, New York, 1976.

    MATH  Google Scholar 

  32. L. Schwartz, Théorie des Distributions, Hermann, Paris, 1950.

    MATH  Google Scholar 

  33. I. Stakgold, Boundary Value Problems of Mathematical Physics, Macmillan, New York, 1968.

    MATH  Google Scholar 

  34. E. Stein, Singular Integrals and Differentiability Properties of Functions, Princeton University Press, Princeton, N. J., 1970.

    MATH  Google Scholar 

  35. E. Stein and G. Weiss, Introduction to Fourier Analysis on Euclidean Space, Princeton University Press, Princeton, N. J., 1971.

    MATH  Google Scholar 

  36. M. Taylor, Noncommutative Harmonic Analysis, Math. Surveys and Monographs, No. 22, AMS, Providence, 1986.

    Google Scholar 

  37. F. Treves, Linear Partial Differential Equations with Constant Coefficients, Gordon and Breach, New York, 1966.

    MATH  Google Scholar 

  38. G. Watson, A Treatise on the Theory of Bessel Functions, Cambridge University Press, Cambridge, 1944.

    MATH  Google Scholar 

  39. E. Whittaker and G. Watson, A Course of Modern Analysis, Cambridge University Press, Cambridge, 1927.

    MATH  Google Scholar 

  40. N. Wiener, Tauberian theorems, Ann. Math. 33(1932), 1–100.

    Google Scholar 

  41. K. Yosida, Functional Analysis, Springer, New York, 1965.

    Book  Google Scholar 

  42. A. Zygmund, Trigonometrical Series, Cambridge University Press, Cambridge, 1969.

    MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Michael E. Taylor .

A The mighty Gaussian and the sublime gamma function

A The mighty Gaussian and the sublime gamma function

The Gaussian function \({e}^{-\vert x{\vert }^{2} }\) on \({\mathbb{R}}^{n}\) is an object whose study yields many wonderful identities. We will use the identity

$${\int \limits _{{\mathbb{R}}^{n}}}{e}^{-\vert x{\vert }^{2} }dx = {\pi }^{n/2},$$
(A.1)

which was established in (3.18), to compute the area An − 1 of the unit sphere Sn − 1 in \({\mathbb{R}}^{n}\). This computation will bring in Euler’s gamma function, and other results will flow from this. Switching to polar coordinates for the right side of (A.1), we have

$$\begin{array}{rlrlrl}{\pi }^{n/2}& = {A}_{ n-1}{\int \nolimits\nolimits }_{0}^{\infty }{e}^{-{r}^{2} }\ {r}^{n-1}\ dr& &\cr & =\frac{1} {2}{A}_{n-1}{\int \nolimits\nolimits }_{0}^{\infty }{e}^{-t}\ {t}^{n/2-1}\mathit{dt}&\cr & =\frac{1} {2}{A}_{n-1}\Gamma {\Bigl (\frac{n} {2}\Bigr)},&\cr\end{array}$$
(A.2)

where the gamma function is defined by

$$\Gamma (z) ={\int \nolimits\nolimits }_{0}^{\infty }{e}^{-t}\ {t}^{z-1}\mathit{dt},$$
(A.3)

for Re z > 0. Thus we have the formula

$${A}_{n-1} =\frac{2{\pi }^{n/2}} {\Gamma\left (\frac{1} {2}n\right)}.$$
(A.4)

To be satisfied with this, we need an explicit evaluation of Γ(n∕2). This can be obtained from Γ(1∕2) and Γ(1) via the following identity:

$$\begin{array}{rlrlrl}\Gamma (z + 1)& ={\int \nolimits\nolimits }_{0}^{\infty }{e}^{-t}\ {t}^{z}\ \mathit{dt}& &\cr & = -{\int \nolimits\nolimits }_{0}^{\infty }\frac{d} {dt}\left ({e}^{-t}\right)\ {t}^{z}\ \mathit{dt}&\cr & = z\Gamma (z),&\cr\end{array}$$
(A.5)

for Re z > 0, where we used integration by parts. The definition (A.3) clearly gives

$$\Gamma (1) = 1.$$
(A.6)

Thus, for any integer k ≥ 1,

$$\Gamma (k) = (k - 1)\Gamma (k - 1) =\cdots = (k - 1)!\,.$$
(A.7)

Note that, for n = 2, we have A1 = 2π∕Γ(1), so (A.6) agrees with the fact that the circumference of the unit circle is 2π (which, of course, figured into the proof of (3.18), via (3.20). In case n = 1, we have A0 = 2, which by (A.4) is equal to 2π1∕2Γ(1∕2), so

$$\Gamma\left (\frac{1} {2}\right) = {\pi }^{1/2}.$$
(A.8)

Again using (A.5), we see that, when k ≥ 1 is an integer,

$$\begin{array}{rlrlrl}\Gamma\left (k +\frac{1} {2}\right)& =\left (k -\frac{1} {2}\right)\Gamma\left (k -\frac{1} {2}\right) =\cdots & &\cr & =\left (k -\frac{1} {2}\right)\left (k -\frac{3} {2}\right)\cdots\left (\frac{1} {2}\right)\Gamma\left (\frac{1} {2}\right)&\cr & = {\pi }^{1/2}\left (k -\frac{1} {2}\right)\left (k -\frac{3} {2}\right)\cdots\left (\frac{1} {2}\right).&\cr\end{array}$$
(A.9)

In particular, Γ(3∕2) = (1∕2)Γ(1∕2) = π1∕2∕2, so A2 = 2π3∕2∕(π1∕2∕2) = 4π,which agrees with the well known formula for the area of the unit sphere in \({\mathbb{R}}^{3}\).

Note that while Γ(z) defined by (A.3) is a priori holomorphic for Re z positive, the equation (A.5) shows that Γ(z) has a meromorphic extension to the entire complex plane, with simple poles at \(z = 0,-1,-2,\ldots\,\). It turns out that Γ(z) has no zeros, so 1∕Γ(z) is an entire analytic function. This is a consequence of the identity

$$\Gamma (z)\Gamma (1 - z) =\frac{\pi } {\sin\pi z},$$
(A.10)

which we now establish. From (A.4) we have (for 0 < Re z < 1)

$$\begin{array}{rlrlrl}\Gamma (z)\Gamma (1 - z)& ={\int \nolimits\nolimits }_{0}^{\infty }{\int \nolimits\nolimits }_{0}^{\infty }{e}^{-(s+t)}{s}^{-z}{t}^{z-1}\mathit{ds}\ \mathit{dt}& &\cr & ={\int \nolimits\nolimits }_{0}^{\infty }{\int \nolimits\nolimits }_{0}^{\infty }{e}^{-u}{v}^{z-1}{(1 + v)}^{-1}\mathit{du}\ \mathit{dv}&\cr & ={\int \nolimits\nolimits }_{0}^{\infty }{(1 + v)}^{-1}{v}^{z-1}\ \mathit{dv},&\cr\end{array}$$
(A.11)

where we have used the change of variables u = s + t, v = ts. With v = ex, the last integral is

$${\int \nolimits\nolimits }_{-\infty }^{\infty }{\left (1 + {e}^{x}\right)}^{-1}\ {e}^{\mathit{xz}}\mathit{dx},$$
(A.12)

which is holomorphic for 0 < Re z < 1, and we want to show that it is equal to the right side of (A.10) on this strip. It suffices to prove identity on the line \(z = 1/2 + i\xi, \xi \in\mathbb{R}\); then (A.12) is equal to the Fourier integral

$${\int \nolimits\nolimits }_{-\infty }^{\infty }{\left (2\cosh\frac{x} {2}\right)}^{-1}\ {e}^{\mathit{ix}\xi }\mathit{dx}.$$
(A.13)

To evaluate this, shift the contour of integration from the real line to the line Im x = −2π. There is a pole of the integrand at x = − πi, and we have (A.13) equal to

$$-{\int \nolimits\nolimits }_{-\infty }^{\infty }{\left (2\cosh\frac{x} {2}\right)}^{-1}{e}^{2\pi\xi }{e}^{ix\xi }\mathit{dx} -\text{ Residue}\cdot (2\pi i).$$
(A.14)

Consequently, (A.13) is equal to

$$-2\pi i\frac{\text{ Residue}} {1 + {e}^{2\pi\xi }} =\frac{\pi } {\cosh\pi\xi },$$
(A.15)

and since π/ sinπ(1∕2 + iξ) = π∕coshπξ, the demonstration of (A.10) is complete.

The integral (A.3) and also the last integral in (A.11) are special cases of the Mellin transform:

$$\mathcal{M}f(z) ={\int \nolimits\nolimits }_{0}^{\infty }f(t){t}^{z-1}\mathit{dt}.$$
(A.16)

If we evaluate this on the imaginary axis:

$${\mathcal{M}}^{\#}f(s) ={\int \nolimits\nolimits }_{0}^{\infty }f(t){t}^{is-1}\mathit{dt},$$
(A.17)

given appropriate growth restrictions on f, this is related to the Fourier transform by a change of variable:

$${\mathcal{M}}^{\#}f(s) ={\int \nolimits\nolimits }_{-\infty }^{\infty }f({e}^{x}){e}^{\mathit{isx}}\mathit{dx}.$$
(A.18)

The Fourier inversion formula and Plancherel formula imply

$$f(r) = {(2\pi)}^{-1}{\int \nolimits\nolimits }_{-\infty }^{\infty }({\mathcal{M}}^{\#}f)(s){r}^{-\mathit{is}}\mathit{ds}$$
(A.19)

and

$${\int \nolimits\nolimits }_{-\infty }^{\infty }\vert {\mathcal{M}}^{\#}f(s){\vert }^{2}\mathit{ds} = (2\pi){\int \nolimits\nolimits }_{0}^{\infty }\vert f(r){\vert }^{2}{r}^{-1}\ dr.$$
(A.20)

In some cases, as seen above, one evaluates ℳf(z) on a vertical line other than the imaginary axis, which introduces only a slight wrinkle.

An important identity for the gamma function follows from taking the Mellin transform with respect to y of both sides of the subordination identity

$${e}^{-yA} =\frac{1} {2}y{\pi }^{-1/2}{\int \nolimits\nolimits }_{0}^{\infty }{e}^{-{y}^{2}/4t }{e}^{-t{A}^{2} }{t}^{-3/2}\mathit{dt}$$
(A.21)

(if y > 0, A > 0), established in §5; see (5.22). The Mellin transform of the left side is clearly Γ(z)Az. The Mellin transform of the right side is a double integral, which is readily converted to a product of two integrals, each defining gamma functions. After a few changes of variables, there results the identity

$${\pi }^{1/2}\Gamma (2z) = {2}^{2z-1}\Gamma (z)\Gamma\left (z +\frac{1} {2}\right),$$
(A.22)

known as the duplication formula for the gamma function. In view of the uniqueness of Mellin transforms, following from (A.18) and (A.19), the identity (A.22) conversely implies (A.21). In fact, (A.22) was obtained first (by Legendre) and this argument produces one of the standard proofs of the subordination identity (A.21).

There is one further identity, which, together with (A.5), (A.10), and (A.22), completes the list of the basic elementary identities for the gamma function. Namely, if the beta function is defined by

$$B(x,y) ={\int \nolimits\nolimits }_{0}^{1}{s}^{x-1}{(1 - s)}^{y-1}\mathit{ds} ={\int \nolimits\nolimits }_{0}^{1}{(1 + u)}^{-x-y}{u}^{x-1}\mathit{du}$$
(A.23)

(with u = s∕(1 − s)), then

$$B(x,y) =\frac{\Gamma (x)\Gamma (y)} {\Gamma (x + y)}.$$
(A.24)

To prove this, note that since

$$\Gamma (z){p}^{-z} ={\int \nolimits\nolimits }_{0}^{\infty }{e}^{-pt}{t}^{z-1}\mathit{dt},$$
(A.25)

we have

$${(1 + u)}^{-x-y} =\frac{1} {\Gamma (x + y)}{\int \nolimits\nolimits }_{0}^{\infty }{e}^{-(1+u)t}{t}^{x+y-1}\mathit{dt},$$

so

$$\begin{array}{rlrlrl}B(x,y)& =\frac{1} {\Gamma (x + y)}{\int \nolimits\nolimits }_{0}^{\infty }{e}^{-t}{t}^{x+y-1}{\int \nolimits\nolimits }_{0}^{\infty }{e}^{-ut}{u}^{x-1}\mathit{du}\ \mathit{dt}& &\cr & =\frac{\Gamma (x)} {\Gamma (x + y)}{\int \nolimits\nolimits }_{0}^{\infty }{e}^{-t}{t}^{y-1}\ \mathit{dt}&\cr & =\frac{\Gamma (x)\Gamma (y)} {\Gamma (x + y)},&\cr\end{array}$$

as asserted.

The four basic identities proved above are the workhorses for most applications involving gamma functions, but fundamental insight is provided by the identities (A.27) and (A.31) below. First, since 0 ≤ et − (1 − tn)nett2n for 0 ≤ tn, we have, for Re z > 0,

$$\begin{array}{rlrlrl}\Gamma (z)& ={\int \nolimits\nolimits }_{0}^{\infty }{e}^{-t}{t}^{z-1}\ \mathit{dt}& &\cr & {=\lim \limits_{n\rightarrow\infty }}{\int \nolimits\nolimits }_{0}^{n}{\Bigl (1 -{\frac{t} {n}}\Bigr)}^{n}{t}^{z-1}\ \mathit{dt}&\cr & {=\lim \limits_{n\rightarrow\infty }}{n}^{z}{\int \nolimits\nolimits }_{0}^{1}{(1 - s)}^{n}{s}^{z-1}\mathit{ds}.&\cr\end{array}$$
(A.26)

Repeatedly integrating by parts gives

$$\Gamma (z) {=\lim \limits_{n\rightarrow\infty }}{n}^{z}\frac{n(n - 1)\cdots 1} {z(z + 1)\cdots (z + n - 1)}{\int \nolimits\nolimits }_{0}^{1}{s}^{z+n-1}\mathit{ds},$$

which yields the following result of Euler:

$$\Gamma (z) {=\lim \limits_{n\rightarrow\infty }}{n}^{z}\frac{1\cdot 2\cdots n} {z(z + 1)\cdots (z + n)}.$$
(A.27)

Using the identity (A.5), analytically continuing Γ(z), we have (A.27) for all z, other than \(0,-1,-2,\ldots \). We can rewrite (A.27) as

$$\Gamma (z) {=\lim \limits_{n\rightarrow\infty }}{n}^{z}\ {z}^{-1}{(1 + z)}^{-1}{\left (1 +\frac{z} {2}\right)}^{-1}\cdots {\left (1 +\frac{z} {n}\right)}^{-1}.$$
(A.28)

If we denote by γ Euler’s constant:

$$\gamma\ {=\lim \limits_{n\rightarrow\infty }}\left (1 +\frac{1} {2} +\cdots +\frac{1} {n} -\log n\right),$$
(A.29)

then (A.28) is equivalent to

$$\Gamma (z) {=\lim \limits_{n\rightarrow\infty }}{e}^{-\gamma z}{e}^{z(1+1/2+\cdots +1/n)}{z}^{-1}{(1 + z)}^{-1}{\left (1 +\frac{z} {2}\right)}^{-1}\cdots {\left (1 +\frac{z} {n}\right)}^{-1},$$
(A.30)

that is, to the Euler product expansion

$$\frac{1} {\Gamma (z)} = z\ {e}^{\gamma z}{\prod \limits_{n=1}^{\infty }}\left (1 +\frac{z} {n}\right){e}^{-z/n}.$$
(A.31)

It follows that the entire analytic function 1∕Γ(z)Γ(− z) has the product expansion

$$\frac{1} {\Gamma (z)\Gamma (-z)} = -{z}^{2}{\prod \limits_{n=1}^{\infty }}\left (1 -\frac{{z}^{2}} {{n}^{2}}\right).$$
(A.32)

Since Γ(1 − z) = − (− z), by virtue of (A.10) this last identity is equivalent to the Euler product expansion

$$\sin\pi z =\pi z{\prod \limits_{n=1}^{\infty }}\left (1 -\frac{{z}^{2}} {{n}^{2}}\right).$$
(A.33)

It is quite easy to deduce the formula (A.5) from the Euler product expansion (A.31). Also, to deduce the duplication formula (A.22) from the Euler product formula is a fairly straightforward exercise.

Finally, we derive Stirling’s formula, for the asymptotic behavior of Γ(z) as z → + . The approach uses the Laplace asymptotic method, which has many other applications. We begin by setting t = sz and then s = ey in the integral formula (A.3), obtaining

$$\begin{array}{rlrlrl}\Gamma (z)& = {z}^{z}{\int \nolimits\nolimits }_{0}^{\infty }{e}^{-z(s-\log s)}{s}^{-1}ds& &\cr & = {z}^{z}{\int \nolimits\nolimits }_{-\infty }^{\infty }{e}^{-z({e}^{y}-y) }\,\mathit{dy}.&\cr\end{array}$$
(A.34)

The last integral is of the form

$${\int \nolimits\nolimits }_{-\infty }^{\infty }{e}^{-z\varphi (y)}\,\mathit{dy},$$
(A.35)

where φ(y) = eyy has a nondegenerate minimum at y = 0; φ(0) = 1, φ(0) = 0, φ(0) = 1. If we write 1=A(y) + B(y), AC0((−2,2)), A(y) = 1 for |y| ≤ 1, then the integral (A.35) is readily seen to be

$${\int \nolimits\nolimits }_{-\infty }^{\infty }A(y){e}^{-z\varphi (y)}dy + O({e}^{-(1+1/e)z}).$$
(A.36)

We can make a smooth change of variable x = ξ(y) such that ξ(y) = y + O(y2), φ(y) = 1 + x2∕2, and the integral in (A.36) becomes

$${e}^{-z}{\int \nolimits\nolimits }_{-\infty }^{\infty }{A}_{ 1}(x){e}^{-z{x}^{2}/2 }\,\mathit{dx},$$
(A.37)

where \({A}_{1}\in{C}_{0}^{\infty }(\mathbb{R}),\ {A}_{1}(0) = 1\), and it is easy to see that, as z → + ,

$${\int \nolimits\nolimits }_{-\infty }^{\infty }{A}_{ 1}(x){e}^{-z{x}^{2}/2 }\,dx\sim {\left (\frac{2\pi } {z}\right)}^{1/2}\left [1 +\frac{{a}_{1}} {z} +\cdots\,\right ].$$
(A.38)

In fact, if z = 1∕2t, then (A.38) is equal to (4πt)1∕2 u(t,0), where u(t, x) solves the heat equation, utuxx = 0, u(0, x) = A1(x). Returning to (A.34), we have Stirling’s formula:

$$\Gamma(z) =\left (\frac{2\pi }/ {z}\right)^{1/2}\ {z}^{z}\ {e}^{-z}\left [1 + O({z}^{-1})\right ].$$
(A.39)

Since n!=Γ(n + 1), we have in particular that

$$n! = {(2\pi n)}^{1/2}\ {n}^{n}\ {e}^{-n}\left [1 + O({n}^{-1})\right ]$$
(A.40)

as n.

Regarding this approach to the Laplace asymptotic method, compare the derivation of the stationary phase method in Appendix B of Chap. 6.

Rights and permissions

Reprints and permissions

Copyright information

© 2011 Springer Science+Business Media, LLC

About this chapter

Cite this chapter

Taylor, M.E. (2011). Fourier Analysis, Distributions, and Constant-Coefficient Linear PDE. In: Partial Differential Equations I. Applied Mathematical Sciences, vol 115. Springer, New York, NY. https://doi.org/10.1007/978-1-4419-7055-8_3

Download citation

Publish with us

Policies and ethics