Method of steepest descent





In mathematics, the method of steepest descent or stationary-phase method or saddle-point method is an extension of Laplace's method for approximating an integral, where one deforms a contour integral in the complex plane to pass near a stationary point (saddle point), in roughly the direction of steepest descent or stationary phase. The saddle-point approximation is used with integrals in the complex plane, whereas Laplace’s method is used with real integrals.


The integral to be estimated is often of the form


Cf(z)eλg(z)dz,{displaystyle int _{C}f(z)e^{lambda g(z)},dz,}{displaystyle int _{C}f(z)e^{lambda g(z)},dz,}

where C is a contour, and λ is large. One version of the method of steepest descent deforms the contour of integration C into a new path integration C′ so that the following conditions hold:




  1. C′ passes through one or more zeros of the derivative g′(z),

  2. the imaginary part of g(z) is constant on C′.


The method of steepest descent was first published by Debye (1909), who used it to estimate Bessel functions and pointed out that it occurred in the unpublished note Riemann (1863) about hypergeometric functions. The contour of steepest descent has a minimax property, see Fedoryuk (2001). Siegel (1932) described some other unpublished notes of Riemann, where he used this method to derive the Riemann–Siegel formula.




Contents






  • 1 A simple estimate[1]


  • 2 The case of a single non-degenerate saddle point


    • 2.1 Basic notions and notation


    • 2.2 Complex Morse lemma


    • 2.3 The asymptotic expansion in the case of a single non-degenerate saddle point




  • 3 The case of multiple non-degenerate saddle points


  • 4 The other cases


  • 5 Extensions and generalizations


  • 6 See also


  • 7 Notes


  • 8 References





A simple estimate[1]


Let f, S : CnC and CCn. If


M=supx∈Cℜ(S(x))<∞,{displaystyle M=sup _{xin C}Re (S(x))<infty ,} M = sup_{x in C} Re (S(x)) < infty,

where (⋅){displaystyle Re (cdot )}Re (cdot) denotes the real part, and there exists a positive real number λ0 such that


C|f(x)eλ0S(x)|dx<∞,{displaystyle int _{C}left|f(x)e^{lambda _{0}S(x)}right|dx<infty ,}int_{C} left| f(x) e^{lambda_0 S(x)}  right| dx < infty,

then the following estimate holds:


|∫Cf(x)eλS(x)dx|⩽const⋅M,∀λR,λλ0.{displaystyle left|int _{C}f(x)e^{lambda S(x)}dxright|leqslant {text{const}}cdot e^{lambda M},qquad forall lambda in mathbb {R} ,quad lambda geqslant lambda _{0}.}left| int_{C} f(x) e^{lambda S(x)}  dx right| leqslant text{const}cdot e^{lambda M}, qquad forall lambda in mathbb{R}, quad lambda geqslant lambda_0.



The case of a single non-degenerate saddle point



Basic notions and notation


Let x be a complex n-dimensional vector, and


Sxx″(x)≡(∂2S(x)∂xi∂xj),1⩽i,j⩽n,{displaystyle S''_{xx}(x)equiv left({frac {partial ^{2}S(x)}{partial x_{i}partial x_{j}}}right),qquad 1leqslant i,,jleqslant n,}S''_{xx}(x) equiv left( frac{partial^2 S(x)}{partial x_i partial x_j} right), qquad 1leqslant i,, jleqslant n,

denote the Hessian matrix for a function S(x). If


φ(x)=(φ1(x),φ2(x),…k(x)){displaystyle {boldsymbol {varphi }}(x)=(varphi _{1}(x),varphi _{2}(x),ldots ,varphi _{k}(x))}boldsymbol{varphi}(x) = (varphi_1(x), varphi_2(x), ldots, varphi_k(x))

is a vector function, then its Jacobian matrix is defined as


φx′(x)≡(∂φi(x)∂xj),1⩽i⩽k,1⩽j⩽n.{displaystyle {boldsymbol {varphi }}_{x}'(x)equiv left({frac {partial varphi _{i}(x)}{partial x_{j}}}right),qquad 1leqslant ileqslant k,quad 1leqslant jleqslant n.}boldsymbol{varphi}_x' (x) equiv left( frac{partial varphi_i (x)}{partial x_j} right), qquad 1 leqslant i leqslant k, quad 1 leqslant j leqslant n.

A non-degenerate saddle point, z0Cn, of a holomorphic function S(z) is a critical point of the function (i.e., S(z0) = 0) where the function's Hessian matrix has a non-vanishing determinant (i.e., detSzz″(z0)≠0{displaystyle det S''_{zz}(z^{0})neq 0}det S''_{zz}(z^0) neq 0).


The following is the main tool for constructing the asymptotics of integrals in the case of a non-degenerate saddle point:



Complex Morse lemma


The Morse lemma for real-valued functions generalizes as follows[2] for holomorphic functions: near a non-degenerate saddle point z0 of a holomorphic function S(z), there exist coordinates in terms of which S(z) − S(z0) is exactly quadratic. To make this precise, let S be a holomorphic function with domain WCn, and let z0 in W be a non-degenerate saddle point of S, that is, S(z0) = 0 and detSzz″(z0)≠0{displaystyle det S''_{zz}(z^{0})neq 0}det S''_{zz}(z^0) neq 0. Then there exist neighborhoods UW of z0 and VCn of w = 0, and a bijective holomorphic function φ : VU with φ(0) = z0 such that


w∈V:S(φ(w))=S(z0)+12∑j=1nμjwj2,detφw′(0)=1,{displaystyle forall win V:qquad S({boldsymbol {varphi }}(w))=S(z^{0})+{frac {1}{2}}sum _{j=1}^{n}mu _{j}w_{j}^{2},quad det {boldsymbol {varphi }}_{w}'(0)=1,}forall w in V: qquad S(boldsymbol{varphi}(w)) = S(z^0) + frac{1}{2} sum_{j=1}^n mu_j w_j^2, quad detboldsymbol{varphi}_w'(0) = 1,<br />

Here, the μj are the eigenvalues of the matrix Szz″(z0){displaystyle S_{zz}''(z^{0})}S_{zz}''(z^0).




An illustration of Complex Morse lemma




The asymptotic expansion in the case of a single non-degenerate saddle point


Assume




  1. f (z) and S(z) are holomorphic functions in an open, bounded, and simply connected set ΩxCn such that the Ix = ΩxRn is connected;


  2. (S(z)){displaystyle Re (S(z))}Re(S(z)) has a single maximum: maxz∈Ixℜ(S(z))=ℜ(S(x0)){displaystyle max _{zin I_{x}}Re (S(z))=Re (S(x^{0}))}max_{z in I_x} Re(S(z)) = Re(S(x^0)) for exactly one point x0Ix;


  3. x0 is a non-degenerate saddle point (i.e., S(x0) = 0 and detSxx″(x0)≠0{displaystyle det S''_{xx}(x^{0})neq 0}det S''_{xx}(x^0) neq 0).


Then, the following asymptotic holds



I(λ)≡Ixf(x)eλS(x)dx=(2πλ)n2eλS(x0)(f(x0)+O(λ1))∏j=1n(−μj)−12,λ,{displaystyle I(lambda )equiv int _{I_{x}}f(x)e^{lambda S(x)}dx=left({frac {2pi }{lambda }}right)^{frac {n}{2}}e^{lambda S(x^{0})}left(f(x^{0})+Oleft(lambda ^{-1}right)right)prod _{j=1}^{n}(-mu _{j})^{-{frac {1}{2}}},qquad lambda to infty ,}I(lambda) equiv int_{I_x} f(x) e^{lambda S(x)} dx = left( frac{2pi}{lambda}right)^{frac{n}{2}} e^{lambda S(x^0)} left(f(x^0)+ Oleft(lambda^{-1}right) right) prod_{j=1}^n (-mu_j)^{-frac{1}{2}}, qquad lambda to infty,
(8)


where μj are eigenvalues of the Hessian Sxx″(x0){displaystyle S''_{xx}(x^{0})}S''_{xx}(x^0) and (−μj)−12{displaystyle (-mu _{j})^{-{frac {1}{2}}}}(-mu_j)^{-frac{1}{2}} are defined with arguments


|arg⁡μj|<π4.{displaystyle left|arg {sqrt {-mu _{j}}}right|<{tfrac {pi }{4}}.}left | argsqrt{-mu_j} right| < tfrac{pi}{4}.

(9)

This statement is a special case of more general results presented in Fedoryuk (1987).[4]



Equation (8) can also be written as



I(λ)=(2πλ)n2eλS(x0)(det(−Sxx″(x0)))−12(f(x0)+O(λ1)),{displaystyle I(lambda )=left({frac {2pi }{lambda }}right)^{frac {n}{2}}e^{lambda S(x^{0})}left(det(-S_{xx}''(x^{0}))right)^{-{frac {1}{2}}}left(f(x^{0})+Oleft(lambda ^{-1}right)right),}I(lambda) = left( frac{2pi}{lambda}right)^{frac{n}{2}} e^{lambda S(x^0)} left ( det (-S_{xx}''(x^0)) right )^{-frac{1}{2}} left (f(x^0) + Oleft(lambda^{-1}right) right),
(13)


where the branch of


det(−Sxx″(x0)){displaystyle {sqrt {det left(-S_{xx}''(x^{0})right)}}}sqrt{det left (-S_{xx}''(x^0) right)}

is selected as follows


(det(−Sxx″(x0)))−12=exp⁡(−i Ind(−Sxx″(x0)))∏j=1n|μj|−12,Ind(−Sxx″(x0))=12∑j=1narg⁡(−μj),|arg⁡(−μj)|<π2.{displaystyle {begin{aligned}left(det left(-S_{xx}''(x^{0})right)right)^{-{frac {1}{2}}}&=exp left(-i{text{ Ind}}left(-S_{xx}''(x^{0})right)right)prod _{j=1}^{n}left|mu _{j}right|^{-{frac {1}{2}}},\{text{Ind}}left(-S_{xx}''(x^{0})right)&={tfrac {1}{2}}sum _{j=1}^{n}arg(-mu _{j}),&&|arg(-mu _{j})|<{tfrac {pi }{2}}.end{aligned}}}begin{align}<br />
left (det left (-S_{xx}''(x^0) right ) right)^{-frac{1}{2}} &= expleft( -i text{ Ind} left (- S_{xx}''(x^0) right ) right) prod_{j=1}^n left| mu_j right|^{-frac{1}{2}}, \<br />
text{Ind} left (-S_{xx}''(x^0) right) &= tfrac{1}{2} sum_{j=1}^n arg (-mu_j), && |arg(-mu_j)| < tfrac{pi}{2}.<br />
end{align}

Consider important special cases:


  • If S(x) is real valued for real x and x0 in Rn (aka, the multidimensional Laplace method), then[7]

Ind(−Sxx″(x0))=0.{displaystyle {text{Ind}}left(-S_{xx}''(x^{0})right)=0.}text{Ind} left(-S_{xx}''(x^0) right ) = 0.

  • If S(x) is purely imaginary for real x (i.e., (S(x))=0{displaystyle Re (S(x))=0}Re(S(x)) = 0 for all x in Rn) and x0 in Rn (aka, the multidimensional stationary phase method),[8] then[9]

Ind(−Sxx″(x0))=π4sign Sxx″(x0),{displaystyle {text{Ind}}left(-S_{xx}''(x^{0})right)={frac {pi }{4}}{text{sign }}S_{xx}''(x_{0}),}text{Ind} left (-S_{xx}''(x^0) right ) = frac{pi}{4} text{sign }S_{xx}''(x_0),

where sign Sxx″(x0){displaystyle {text{sign }}S_{xx}''(x_{0})}text{sign }S_{xx}''(x_0) denotes the signature of matrix Sxx″(x0){displaystyle S_{xx}''(x_{0})}S_{xx}''(x_0), which equals to the number of negative eigenvalues minus the number of positive ones. It is noteworthy that in applications of the stationary phase method to the multidimensional WKB approximation in quantum mechanics (as well as in optics), Ind is related to the Maslov index see, e.g., Chaichian & Demichev (2001) and Schulman (2005).


The case of multiple non-degenerate saddle points


If the function S(x) has multiple isolated non-degenerate saddle points, i.e.,


S(x(k))=0,detSxx″(x(k))≠0,x(k)∈Ωx(k),{displaystyle nabla Sleft(x^{(k)}right)=0,quad det S''_{xx}left(x^{(k)}right)neq 0,quad x^{(k)}in Omega _{x}^{(k)},}nabla S left (x^{(k)} right ) = 0, quad det S''_{xx} left (x^{(k)} right ) neq 0, quad x^{(k)} in Omega_x^{(k)},

where


x(k)}k=1K{displaystyle left{Omega _{x}^{(k)}right}_{k=1}^{K}}left { Omega_x^{(k)} right }_{k=1}^K

is an open cover of Ωx, then the calculation of the integral asymptotic is reduced to the case of a single saddle point by employing the partition of unity. The partition of unity allows us to construct a set of continuous functions ρk(x) : Ωx → [0, 1], 1 ≤ kK, such that


k=1Kρk(x)=1,∀x∈Ωx,ρk(x)=0∀x∈Ωx∖Ωx(k).{displaystyle {begin{aligned}sum _{k=1}^{K}rho _{k}(x)&=1,&&forall xin Omega _{x},\rho _{k}(x)&=0&&forall xin Omega _{x}setminus Omega _{x}^{(k)}.end{aligned}}}begin{align} <br />
sum_{k=1}^K rho_k(x) &= 1, && forall x in Omega_x, \<br />
rho_k(x) &= 0 && forall x in Omega_xsetminus Omega_x^{(k)}.<br />
end{align}

Whence,


Ix⊂Ωxf(x)eλS(x)dx≡k=1K∫Ix⊂Ωk(x)f(x)eλS(x)dx.{displaystyle int _{I_{x}subset Omega _{x}}f(x)e^{lambda S(x)}dxequiv sum _{k=1}^{K}int _{I_{x}subset Omega _{x}}rho _{k}(x)f(x)e^{lambda S(x)}dx.}int_{I_x subset Omega_x} f(x) e^{lambda S(x)} dx equiv sum_{k=1}^K int_{I_x subset Omega_x} rho_k(x) f(x) e^{lambda S(x)} dx.

Therefore as λ → ∞ we have:


k=1K∫a neighborhood of x(k)f(x)eλS(x)dx=(2πλ)n2∑k=1KeλS(x(k))(det(−Sxx″(x(k))))−12f(x(k)),{displaystyle sum _{k=1}^{K}int _{{text{a neighborhood of }}x^{(k)}}f(x)e^{lambda S(x)}dx=left({frac {2pi }{lambda }}right)^{frac {n}{2}}sum _{k=1}^{K}e^{lambda Sleft(x^{(k)}right)}left(det left(-S_{xx}''left(x^{(k)}right)right)right)^{-{frac {1}{2}}}fleft(x^{(k)}right),}sum_{k=1}^K  int_{text{a neighborhood of }x^{(k)}} f(x) e^{lambda S(x)} dx = left(frac{2pi}{lambda}right)^{frac{n}{2}} sum_{k=1}^K e^{lambda S left (x^{(k)} right )} left ( det left(-S_{xx}'' left (x^{(k)} right )right) right)^{-frac{1}{2}} f left (x^{(k)} right ),

where equation (13) was utilized at the last stage, and the pre-exponential function f (x) at least must be continuous.



The other cases


When S(z0) = 0 and detSzz″(z0)=0{displaystyle det S''_{zz}(z^{0})=0}det S''_{zz}(z^0) = 0, the point z0Cn is called a degenerate saddle point of a function S(z).


Calculating the asymptotic of


f(x)eλS(x)dx,{displaystyle int f(x)e^{lambda S(x)}dx,} int f(x) e^{lambda S(x)} dx,

when λ → ∞,  f (x) is continuous, and S(z) has a degenerate saddle point, is a very rich problem, whose solution heavily relies on the catastrophe theory. Here, the catastrophe theory replaces the Morse lemma, valid only in the non-degenerate case, to transform the function S(z) into one of the multitude of canonical representations. For further details see, e.g., Poston & Stewart (1978) and Fedoryuk (1987).


Integrals with degenerate saddle points naturally appear in many applications including optical caustics and the multidimensional WKB approximation in quantum mechanics.


The other cases such as, e.g., f (x) and/or S(x) are discontinuous or when an extremum of S(x) lies at the integration region's boundary, require special care (see, e.g., Fedoryuk (1987) and Wong (1989)).



Extensions and generalizations


An extension of the steepest descent method is the so-called nonlinear stationary phase/steepest descent method. Here, instead of integrals, one needs to evaluate asymptotically solutions of Riemann–Hilbert factorization problems.


Given a contour C in the complex sphere, a function f defined on that contour and a special point, say infinity, one seeks a function M holomorphic away from the contour C, with prescribed jump across C, and with a given normalization at infinity. If f and hence M are matrices rather than scalars this is a problem that in general does not admit an explicit solution.


An asymptotic evaluation is then possible along the lines of the linear stationary phase/steepest descent method. The idea is to reduce asymptotically the solution of the given Riemann–Hilbert problem to that of a simpler, explicitly solvable, Riemann–Hilbert problem. Cauchy's theorem is used to justify deformations of the jump contour.


The nonlinear stationary phase was introduced by Deift and Zhou in 1993, based on earlier work of the Russian mathematician Alexander Its. A (properly speaking) nonlinear steepest descent method was introduced by Kamvissis, K. McLaughlin and P. Miller in 2003, based on previous work of Lax, Levermore, Deift, Venakides and Zhou. As in the linear case, steepest descent contours solve a min-max problem. In the nonlinear case they turn out to be "S-curves" (defined in a different context back in the 80s by Stahl, Gonchar and Rakhmanov).


The nonlinear stationary phase/steepest descent method has applications to the theory of soliton equations and integrable models, random matrices and combinatorics.



See also


Pearcey integral



Notes





  1. ^ A modified version of Lemma 2.1.1 on page 56 in Fedoryuk (1987).


  2. ^ Lemma 3.3.2 on page 113 in Fedoryuk (1987)


  3. ^ Poston & Stewart (1978), page 54; see also the comment on page 479 in Wong (1989).


  4. ^ Fedoryuk (1987), pages 417-420.


  5. ^ This conclusion follows from a comparison between the final asymptotic for I0(λ), given by equation (8), and a simple estimate for the discarded integral I1(λ).


  6. ^ This is justified by comparing the integral asymptotic over Rn [see equation (8)] with a simple estimate for the altered part.


  7. ^ See equation (4.4.9) on page 125 in Fedoryuk (1987)


  8. ^ Rigorously speaking, this case cannot be inferred from equation (8) because the second assumption, utilized in the derivation, is violated. To include the discussed case of a purely imaginary phase function, condition (9) should be replaced by |arg⁡μj|⩽π4.{displaystyle left|arg {sqrt {-mu _{j}}}right|leqslant {tfrac {pi }{4}}.} left | argsqrt{-mu_j} right | leqslant tfrac{pi}{4}.


  9. ^ See equation (2.2.6') on page 186 in Fedoryuk (1987)




References




  • Chaichian, M.; Demichev, A. (2001), Path Integrals in Physics Volume 1: Stochastic Process and Quantum Mechanics, Taylor & Francis, p. 174, ISBN 075030801X.mw-parser-output cite.citation{font-style:inherit}.mw-parser-output q{quotes:"""""""'""'"}.mw-parser-output code.cs1-code{color:inherit;background:inherit;border:inherit;padding:inherit}.mw-parser-output .cs1-lock-free a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/6/65/Lock-green.svg/9px-Lock-green.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .cs1-lock-limited a,.mw-parser-output .cs1-lock-registration a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/d/d6/Lock-gray-alt-2.svg/9px-Lock-gray-alt-2.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .cs1-lock-subscription a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/a/aa/Lock-red-alt-2.svg/9px-Lock-red-alt-2.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registration{color:#555}.mw-parser-output .cs1-subscription span,.mw-parser-output .cs1-registration span{border-bottom:1px dotted;cursor:help}.mw-parser-output .cs1-hidden-error{display:none;font-size:100%}.mw-parser-output .cs1-visible-error{font-size:100%}.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registration,.mw-parser-output .cs1-format{font-size:95%}.mw-parser-output .cs1-kern-left,.mw-parser-output .cs1-kern-wl-left{padding-left:0.2em}.mw-parser-output .cs1-kern-right,.mw-parser-output .cs1-kern-wl-right{padding-right:0.2em}


  • Debye, P. (1909), "Näherungsformeln für die Zylinderfunktionen für große Werte des Arguments und unbeschränkt veränderliche Werte des Index", Mathematische Annalen, 67 (4): 535–558, doi:10.1007/BF01450097 English translation in Debye, Peter J. W. (1954), The collected papers of Peter J. W. Debye, Interscience Publishers, Inc., New York, ISBN 978-0-918024-58-9, MR 0063975


  • Deift, P.; Zhou, X. (1993), "A steepest descent method for oscillatory Riemann-Hilbert problems. Asymptotics for the MKdV equation", Ann. of Math., The Annals of Mathematics, Vol. 137, No. 2, 137 (2), pp. 295–368, arXiv:math/9201261, doi:10.2307/2946540, JSTOR 2946540.


  • Erdelyi, A. (1956), Asymptotic Expansions, Dover.


  • Fedoryuk, M V (2001) [1994], "Saddle_point_method", in Hazewinkel, Michiel, Encyclopedia of Mathematics, Springer Science+Business Media B.V. / Kluwer Academic Publishers, ISBN 978-1-55608-010-4.


  • Fedoryuk, M. V. (1987), Asymptotic: Integrals and Series, Nauka, Moscow [in Russian].


  • Kamvissis, S.; McLaughlin, K. T.-R.; Miller, P. (2003), "Semiclassical Soliton Ensembles for the Focusing Nonlinear Schrödinger Equation", Annals of Mathematics Studies, Princeton University Press, 154.


  • Riemann, B. (1863), Sullo svolgimento del quoziente di due serie ipergeometriche in frazione continua infinita (Unpublished note, reproduced in Riemann's collected papers.)


  • Siegel, C. L. (1932), "Über Riemanns Nachlaß zur analytischen Zahlentheorie", Quellen Studien zur Geschichte der Math. Astron. und Phys. Abt. B: Studien 2: 45–80 Reprinted in Gesammelte Abhandlungen, Vol. 1. Berlin: Springer-Verlag, 1966.


  • Poston, T.; Stewart, I. (1978), Catastrophe Theory and Its Applications, Pitman.


  • Schulman, L. S. (2005), "Ch. 17: The Phase of the Semiclassical Amplitude", Techniques and Applications of Path Integration, Dover, ISBN 0486445283


  • Wong, R. (1989), Asymptotic approximations of integrals, Academic Press.




Popular posts from this blog

Shashamane

Carrot

Deprivation index