You are on page 1of 14

INFORMS Journal on Computing informs ®

Vol. 18, No. 4, Fall 2006, pp. 408–421


issn 1091-9856  eissn 1526-5528  06  1804  0408 doi 10.1287/ijoc.1050.0137
© 2006 INFORMS

A Unified Framework for Numerically Inverting


Laplace Transforms
Joseph Abate
900 Hammond Road, Ridgewood, New Jersey 07450-2908, USA

Ward Whitt
Department of Industrial Engineering and Operations Research, Columbia University,
New York, New York 10027-6699, USA, ww2040@columbia.edu

W e introduce and investigate a framework for constructing algorithms to invert Laplace transforms numer-
ically. Given a Laplace transform fˆ of a complex-valued function of a nonnegative real-variable, f , the
function f is approximated by a finite linear combination of the transform values; i.e., we use the inversion
formula  
1 n

f t ≈ fn t ≡ k fˆ k 0 < t < 
t k=0 t
where the weights k and nodes k are complex numbers, which depend on n, but do not depend on the trans-
form fˆ or the time argument t. Many different algorithms can be put into this framework, because it remains
to specify the weights and nodes. We examine three one-dimensional inversion routines in this framework:
the Gaver-Stehfest algorithm, a version of the Fourier-series method with Euler summation, and a version of
the Talbot algorithm, which is based on deforming the contour in the Bromwich inversion integral. We show
that these three building blocks can be combined to produce different algorithms for numerically inverting
two-dimensional Laplace transforms, again all depending on the single parameter n. We show that it can be
advantageous to use different one-dimensional algorithms in the inner and outer loops.
Key words: Laplace transforms; numerical transform inversion; Fourier-series method; Talbot’s method;
Gaver-Stehfest algorithm; multi-precision computing; multidimensional Laplace transforms;
multidimensional transform inversion
History: Accepted by Edward P. C. Kao, Area Editor for Computational Probability and Analysis; received
August 2004; revised December 2004; accepted February 2005.

1. Introduction The flexible framework for numerically inverting


In recent years, numerical-transform inversion has one-dimensional Laplace transforms is also conve-
become recognized as an important technique in oper- nient for developing algorithms to invert multi-
ations research, notably for calculating probability dimensional Laplace transforms. We can combine
distributions in stochastic models. There have now different one-dimensional inversion algorithms to
been many applications; e.g., see the survey by Abate obtain different multi-dimensional inversion algo-
et al. (1999) and the textbook treatment by Kao (1997). rithms. In particular, given three different one-
Over the years, many different algorithms have dimensional inversion algorithms, we directly obtain
been proposed for numerically inverting Laplace nine different two-dimensional inversion algorithms,
transforms; e.g., see the surveys in Abate and Whitt one for each possible combination in the inner and
(1992) and Chapter 19 of Davies (2002), the exten- outer loops. These one-dimensional algorithms can
sive bibliography of Valko and Vojta (2001) and the be combined with hardly any modifications, in some
numerical comparisons by Davies and Martin (1979), cases with none at all. Moreover, we show that it can
Narayanan and Beskos (1982) and Duffy (1993). In be advantageous from the point of view of algorithm
contrast to the usual approach, in this paper we do efficiency to combine two different one-dimensional
not focus on a particular procedure, but instead intro- inversion algorithms in the two-dimensional inver-
duce and investigate a framework that can encompass sion algorithm.
a wide range of procedures. The flexible framework The framework for one-dimensional inversion was
opens the way to further work, e.g., performing opti- originally introduced in several short papers by
mization in order to choose the best set of parameters Zakian (1969, 1970, 1973), but it does not seem to
and thus the best procedure, by various criteria, for have received much attention, if any. Moreover, in
various classes of functions. those papers, Zakian did not mention the advantage
408
Abate and Whitt: A Unified Framework for Numerically Inverting Laplace Transforms
INFORMS Journal on Computing 18(4), pp. 408–421, © 2006 INFORMS 409

of having a flexible framework, encompassing sev- of the approximation will depend on the transform fˆ
eral different procedures. Indeed, Zakian proposed a and the time argument t, but there is the potential
specific procedure, as we will explain in Section 2. for great efficiency. Moreover, the same framework
The major contribution here, we believe, is suggest- applies to multiple procedures as well, so that inde-
ing the more general flexible framework, encompass- pendent accuracy checks are readily available too.
ing several different procedures, and showing how In most applications, f will be real-valued; then we
that can be exploited in multidimensional inversion approximate by the real part, i.e.,
algorithms.   
To demonstrate the potential of our proposed flexi- 1 n
ˆ k
Ref t ≈ Refn t ≡ Re k f
ble framework, we look at three specific widely used t k=0 t
approaches in this framework, namely, (i) Fourier    
1 n

series expansions with Euler summation, (ii) com- = Rek Re fˆ k
binations of Gaver functionals, and (iii) deforma- t k=0 t
tion of the contour in the Bromwich integral. These   

methods are well known as the Euler algorithm (the − Imk Im fˆ k  (3)
t
Fourier-series method with Euler summation), e.g.,
see Dubner and Abate (1968), Abate and Whitt (1992, The case of complex-valued functions arises naturally
1995), Choudhury et al. (1994a), O’Cinneide (1997), when we consider inner loops in the inversion of mul-
Section 19.6 of Davies (2002), and Sakurai (2004); the tidimensional transforms; see Section 3.
Gaver-Stehfest algorithm, e.g., see Gaver (1966), Stehfest For any given n, and any given set of approximat-
(1970), Section 8 of Abate and Whitt (1992), Section ing parameters k k 1 ≤ k ≤ n, the function fn on
19.2 of Davies (2002), and Valko and Abate (2004); and the right in (2) serves as an approximation of the func-
Talbot’s method, e.g., see Talbot (1979), Murli and Riz- tion f , which we intend to apply for all t that are
zardi (1990), Evans (1993), Evans and Chung (2000), continuity points of f with 0 < t < . It is understood
Section 19.8 of Davies (2002), and Abate and Valko that there may be numerical difficulties if t is either
(2004). very small or very large. In those exceptional cases
Here is the idea: given a Laplace transform fˆ of it is often preferable to apply asymptotics as t → 0
a complex-valued function f of a nonnegative real- or t → , e.g., from the initial-value and final-value
variable, theorems (Doetsch 1974); e.g., as in Abate and Whitt
  (1997). Alternatively, it may be desirable to scale the
fˆs ≡ e−st f t dt (1)
0 function, as in Choudhury and Whitt (1997).
the function f is represented approximately by a finite For any specific t, we can increase n in order to
linear combination of the transform values, via obtain better accuracy. However, greater system preci-
  sion will typically be required to perform the calcula-
1 n
 tion as n increases. As in Abate and Valko (2004), we
f t ≈ fn t ≡ k fˆ k t > 0 (2)
t k=0 t suggest exploiting multi-precision software, which is
now widely available through computer algebra sys-
where the nodes k and weights k (i.e., the interior tems such as Mathematica and Maple in order to obtain
and exterior scaling constants) are complex numbers, the required precision; e.g., see Graf (2004). Such high
which depend on n, but not on the transform fˆ or precision is also an integral part of UBASIC, which
the time argument t. We assume that the Laplace the first author has been using for more than twenty
transform fˆs in (1) is well defined and analytic for years.
Res > 0, where Res is the real part of the complex In fact, in this paper we assume that multi-precision
variable
√ s ≡ u + iv; i.e., Res = u for s = u + iv with software is being used, so we do not consider spe-
i = −1 and u and v real numbers. The associated cial measures to control roundoff error within lim-
imaginary part of s is Ims = v. (We have assumed ited precision (e.g., standard double precision). The
that the “abscissa of convergence” is less than or equal framework may also be useful with limited preci-
to 0, which is natural for a large class of applica- sion, provided appropriate measures are taken to con-
tions, but that can be generalized. Indeed, it is without trol roundoff error, but some procedures, such as
loss of generality, because we can make a change of the Gaver-Stehfest algorithm, usually require high
variables.) precision.
Clearly, there are advantages to having the nodes For one of the specific inversion routines we con-
k and weights k in the representation of fn t in (2) sider here (the Euler algorithm), the required pre-
be independent of both the transform fˆ and the time cision as a function of n was identified on p. 272
argument t. As a consequence, the procedure can be of Abate et al. (1999). For the other two specific
efficiently applied to multiple transforms and multi- inversion routines we consider here, Abate and Valko
ple time points. Of course, in general, the accuracy (2004) identified the required precision as a function
Abate and Whitt: A Unified Framework for Numerically Inverting Laplace Transforms
410 INFORMS Journal on Computing 18(4), pp. 408–421, © 2006 INFORMS

of n for a large class of transforms. (The full story real variables, f t1 t2 , the corresponding two-dimen-
is rather complicated; see Abate and Valko 2004 sional inversion formula is
for more details.) Thus in our implementations of
the three basic methods we obtain “automatic algo- f t1 t2  ≈ fn1 n2 t1 t2 
rithms”: given the accuracy we want for f t (by  
1  n1
n2
k1 k2
which we mean relative accuracy, measured in signif- ≡ k1 k2 f˜ (5)
icant digits; the number of significant digits is defined t1 t2 k =0 k =0
t 1 t2
1 2

as − log10 relative error; see (34)), a formula specifies


where n1 and n2 are the parameters, k1 and k2 are the
the appropriate parameter n; then, given n, the pro-
indices, k1 and k2 are the weights, and k1 and k2
gram determines the required precision for the com-
are the nodes for the outer and inner loops, respec-
putation, using another formula, so that the desired
accuracy for f t is realized. Of course, if there are dif- tively. We can calculate the approximant fn1 n2 t1 t2 
ficulties, then the parameter n or the system precision at any argument t1 t2  provided that we can evaluate
can be increased. the double transform f˜s1 s2  at any argument s1 s2 .
Different specific algorithms are obtained by intro- As in (2) and (3), we use the real part if f t1 t2  is
ducing different families of approximating param- real-valued.
eters k k 0 ≤ k ≤ n. Indeed, Stehfest (1970) Given the three specific one-dimensional algo-
originally expressed the Gaver-Stehfest method in this rithms we consider in the framework (2), namely,
form. We will cast the Fourier-series method with Gaver (), Euler (), and Talbot ( ), we can com-
Euler summation and Talbot’s method in this form as bine them in every possible combination to obtain
well. It turns out that other methods also are already nine two-dimensional algorithms. In particular, it is
in this form. In particular, the Padé method used by not necessary to use the same routine in both the
Zakian (1969, 1970) and Vlach (1969) and the Gaus- inner and outer loop. Indeed, we will show that it can
sian quadrature method of Piessens (1969, 1971) and be advantageous to combine two different one-dimensional
Wellekens (1970) are in this form, but we will not routines. We are unaware of this ever having been
examine them in detail here. done before.
On the other hand, there also are methods that are In this paper, we examine these nine two-dimen-
not, at least directly, in the form of (2). Among these sional algorithms, using the notation  , where the
is the popular Laguerre or Weeks method; e.g., see first operator  applies to the outer loop, while the
Abate et al. (1998) and Section 19.5 of Davies (2002). second operator  applies to the inner loop. We dis-
There we have cuss the performance of these two-dimensional rou-
tines, illustrating with numerical examples.

n
f t ≈ ci li bt (4) Here is how the rest of this paper is organized:
i=0 we start in Section 2 by explaining the unified
framework. Then in Section 3 we elaborate on the
where li t, i ≥ 0, are the standard Laguerre functions,
extension of the framework to two dimensions. In
which form an orthonormal basis for the function
the next three sections we introduce specific algo-
space L2 0  R, while the weights ci and the scale
rithms in the framework. In Sections 4, 5, and 6 we
factor b depend on the transform fˆ.
briefly specify our versions of the one-dimensional
A main purpose for the unified framework is to
Gaver-Stehfest, Fourier-series (with Euler summa-
develop new algorithms for multidimensional trans-
tion), and Talbot algorithms in the unified frame-
form inversion. So far, relatively little attention has
work. Then in Section 7 we discuss the perfor-
been given to inversion of multidimensional Laplace
mance of these one-dimensional algorithms. In Sec-
transforms; e.g., see Choudhury et al. (1994a), Abate
tion 8 we show how the three one-dimensional inver-
et al. (1998), Hwang and Lu (1999), and references
sion algorithms can be combined to produce specific
therein. The Fourier-series multidimensional inver-
two-dimensional algorithms. In Section 9 we present
sion routines in Choudhury et al. (1994a) have been
numerical examples evaluating the performance of
applied to tackle challenging performance-analysis
the two-dimensional inversion algorithms. In Section
problems in the steady-state analysis of stochastic net-
10 we draw conclusions.
works and the time-dependent analysis of queues
with time-varying rates in Choudhury et al. (1994b,
1995a, b, 1997). Choudhury et al. (1997) includes an 2. The Unified Framework
example of a 21-dimensional inversion. In this section we provide motivation for the gen-
It is significant that the simple framework in (2) eral inversion formula (2), explaining how it was
extends easily to multidimensional transforms. For originally derived. We present two derivations: first,
any two-dimensional transform f˜s1 s2  of a two- one based on the Bromwich inversion integral due to
dimensional complex-valued function of nonnegative Wellekens (1970) and, second, the original one based
Abate and Whitt: A Unified Framework for Numerically Inverting Laplace Transforms
INFORMS Journal on Computing 18(4), pp. 408–421, © 2006 INFORMS 411

on approximating delta functions due to Zakian (1969, Rational Approximations for the Exponential
1970, 1973). However, we will not use those spe- Function
cific inversion routines later in the paper. We start by To achieve the unified framework in (2) from (9),
observing that it is natural to have t −1 appear both Wellekens (1970) approximated the exponential func-
inside and outside the transform fˆ in the inversion tion ez in the integrand of (9) by a rational function.
formula for f t in (2). Indeed, both of the earlier approaches exploit rational
approximations for the exponential function, so we
Initial-Value and Final-Value Theorems digress to discuss that intermediate step now.
One way to see the role of t −1 is to consider appropri- Writing the rational function in a partial-fraction rep-
ate Abelian and Tauberian theorems, in particular, the resentation (e.g., see p. 76 of Doetsch 1974), we have
familiar initial-value and final-value theorems; e.g.,
see Bingham et al. (1987), Doetsch (1974), or Feller 
n
k
(1971). In simple form, the initial-value theorem states ez ≈ (10)
 −z
k=0 k
that, when the two limits both exist,
 
1 ˆ 1 where z is a complex variable and k and k are com-
lim f t = lim f (6) plex numbers.
t↓0 t↓0 t t
We will impose regularity conditions on the com-
while the final-value theorem states that, when the two plex numbers k and k appearing in (10): First, we
limits both exist, will assume that the parameters k are distinct, imply-
  ing that the rational approximant in (10) has n + 1
1 1
lim f t = lim fˆ  (7) poles. We will require that the approximant be ana-
t↑ t↑ t t
lytic for Rez > 0. Consequently, we assume that
Moreover, other refined asymptotics agree when the Rek  < 0 for all k. We also assume that the pairs of
transform is scaled in that way, e.g., see Heavyside’s complex numbers k k  occur in complex-conjugate
theorem, Theorem 37.1 of Doetsch (1974). pairs. (If s ≡ u + iv, then its complex conjugate is s̄ ≡
u − iv.) That is, if either k or k has a non-zero imag-
Starting from the Bromwich Inversion Integral inary part, then we also have the pair  k 
 k  among
We now start to review Wellekens’ (1970) argument, the n terms in the sum (10). Since
which begins with the Bromwich inversion integral.
Since the Euler and Talbot algorithms also can be k k
derived from the Bromwich inversion integral, this +
k − z k − z
argument explains why they too fit in the frame-
work (2). is real for all real z, that condition guarantees that
Suppose that f is a real-valued function of a non- the approximating rational function is real-valued for
negative real variable, which is continuous at t. The all real z. Consequently, we can alternatively express
Bromwich inversion integral expresses f t as the con- (10) as
tour integral n  
z 1 k 
k
e ≈ + (11)
1  ˆ 2 k=0 k − z  k − z
f t = f sest ds t > 0 (8)
2i C where Rek  < 0 for all k.
where the contour C extends from c − i to c + i for Of course, we want to choose the rational function
c > 0, falling to the right of all singularities of fˆ, under in (10) or (11) so that we obtain a good approxima-
the usual regularity conditions; see Theorem 24.4 of tion. A classic way to do that is to choose the complex
Doetsch (1974). numbers k and k so as to match the MacLaurin
Following Wellekens (1970), we make the change of series of the two functions as far as possible. That pro-
variables z = st and rewrite the contour integral as cedure is effective; indeed, that procedure is the classi-
cal method of Padé approximation; moreover, the expo-
1  ˆ
f t = f z/tez dz t > 0 (9) nential function was one of the first functions to be
2it C  treated; see the Preface and Section 1.2 of Baker and
where C  is the same contour as a function of z. Graves-Morris (1996). It should thus not be surprising
By that step alone, we see how t −1 appears both in that Zakian (1969, 1970, 1973) and Wellekens (1970)
the multiplicative constant and the argument of the exploited Padé approximation when they approxi-
transform fˆ. Indeed, almost any method of numer- mated the exponential function, but of course there
ical integration (approximate quadrature) applied to are many other possible rational approximations for
the integral in (9) produces the representation (2); see the exponential function, with merits depending on
Davis and Rabinowitz (1984). the context.
Abate and Whitt: A Unified Framework for Numerically Inverting Laplace Transforms
412 INFORMS Journal on Computing 18(4), pp. 408–421, © 2006 INFORMS

The Rest of Wellekens’ Argument each continuity point t of f , we approximate f t by


Continuing to follow Wellekens (1970), we achieve the the associated integrals with respect to !n x/t − 1; i.e.,
unified framework in (2) from (9) by approximating combining (14) and (15), we use the approximation
the exponential function ez in the integrand of (9) by
a rational function, using (10) or (11). For example, 1 T
f t ≈ fn t ≡ f x!n x/t − 1 dx 0<t<T
using (10), we get t 0
1  ˆ n
k 1 T  n

fn t ≈ f z/t dz = f x k e−k x/t dx


2it C   −z t 0 k=0
k=0 k
1
n 
1  ˆ
T
1 n
k = k f xe−k x/t dx
= f z/t dz t
t k=0 2i C  k − z k=0 0

  1 n  
1 n
ˆ k → k f xe−k x/t dx
= f t > 0 (12) t k=0
t k=0 k t 0
 
1 n
ˆ k
using the Cauchy integral formula in the last step, = k f (16)
e.g., p. 9 of Davies (2002). t k=0 t
It still remains to choose the specific rational approx-
imation for the exponential function. Wellekens (1970) where we let T →  in the last step, with the limit
shows that the classic Padé approximation arises being justified by standard assumptions on f guaran-
when we perform Gaussian quadrature to the inte- teeing that the Laplace transform is well defined. We
gral (9). Thus there is additional justification for Padé see that (16) is of the same form as (2).
approximation. The linearity in formula (2) implies that the inver-
sion can be applied either to the complex-valued func-
Zakian’s Original Argument tion f or to its real and imaginary parts: Suppose that
Zakian originally obtained (2) in a very different way, f is a complex-valued function and let f 1 ≡ Ref 
in particular, by approximating a delta function by a and f 2 ≡ Imf  be the real and imaginary parts of f .
finite linear combination of exponential functions. Let Then
!x/t − 1 be the scaled delta function that attaches
mass t to the point t, for each t > 0; i.e., it satisfies the f t = f 1 t+if 2 t fˆs = fˆ1 s+ifˆ2 s (17)
properties
 T and
!x/t − 1 dx = t and !x/t − 1 = 0  
0 1 n
ˆ k
fn t ≡ k f
for x = t 0 < t < T  (13) t k=0 t
    
More precisely, if f is an integrable complex-valued 1 n
 
= k fˆ1 k + ifˆ2 k
function of a nonnegative real variable that is contin- t k=0 t t
uous at t, then    
1 n
1 k 1 n
2 k
1 T = f ˆ +i f ˆ
f t = f x!x/t − 1 dx 0 < t < T  (14) t k=0 k t t k=0 k t
t 0      
1 n
ˆ1 k 1 n
ˆ2 k
Note that ! actually is a function of two variables, = Re k f +i Re k f
x and t. t k=0 t t k=0 t
The idea, then, is to approximate the scaled delta ≡ fn1 t + ifn2 t (18)
function !x/t − 1 by a finite linear combination of
exponential functions; i.e., It remains to show that the approximants !n in (15)

n can indeed serve as good approximants for the scaled
!x/t − 1 ≈ !n x/t − 1 ≡ k e−k x/t (15) delta function !. At first, it may not be obvious that
k=0
the delta function admits a reasonable approximation
where k and k are complex numbers. It may not be of the form (15). Indeed, it is obvious that the approx-
immediately clear that this is a good thing to do, but imation !n cannot be a good approximation for the
we explain later. delta function ! uniformly in the neighborhood of its
Now we consider an integrable function f that we discontinuity, since the approximant is a continuous
want to evaluate, again assumed to be a complex- function, but nevertheless we can hope to approxi-
valued function of a nonnegative real variable. For mate f t at points t that are continuity points of f .
Abate and Whitt: A Unified Framework for Numerically Inverting Laplace Transforms
INFORMS Journal on Computing 18(4), pp. 408–421, © 2006 INFORMS 413

As Zakian (1970) observed, the possibility of good interest tj . Let  and  be the vectors of weights
approximations is evident if we take Laplace trans- and nodes, i.e.,  ≡ 0    n  and  ≡ 0    n .
forms: The Laplace transform of the scaled delta func- Then our object might be to choose the vectors  and
tion is  in order to minimize the squared error. That is, we
  might perform the minimization
ˆ t ≡
!s e−sx !x/t − 1 dx = te−ts (19) m1 m2   2
0   1 n

min fi tj  − k fˆi k (21)
while the Laplace transform of the general nth approx-  
i=1 j=1
tj k=0 tj
imant is
where the decision variables  and  are vectors of
  n
k
!ˆ n s t ≡ e−sx !n x/t − 1 dx =  (20) complex numbers. In the optimization (21), the test
0 k=0
s + k /t functions fi and their transforms fˆi are presumed to
be known. We might also put further constraints upon
Assuming that the nodes k are distinct for 0 ≤ k ≤ n, the variables  and . For example, paralleling the
we see that the Laplace transform of the approxi- Gaver-Stehfest procedure, we could require that they
mant is a proper rational function with poles at k /t, be real numbers. Finally, given the vectors  and ,
expressed in a partial-fraction representation with we have a new algorithm in the unified framework (2)
partial fraction coefficients k . to apply to new transforms of interest. This seems to
At this point we exploit uniqueness and continu- be an interesting direction of research, but we do not
ity theorems for Laplace transforms, e.g., see Doetsch pursue it here; see Audis and Whitt (2006).
(1974) and Chapter 13 of Feller (1971), to justify
characterizing functions and measures in terms of
Laplace transforms. Thus, assuming that we can work
3. Two-Dimensional Inversion
with Laplace transforms, in order to approximate Algorithms
the scaled delta function !x/t − 1 by its approxi- Starting from a function of two variables, f t1 t2 ,
mant !n x/t − 1 in (15), it suffices to approximate we can construct the Laplace transform in a two-
the exponential function te−ts in (19) by the ratio- step process, applying the one-dimensional construc-
nal functions in (20). As observed by Zakian (1970), tion twice: first, consider t2 a constant, and apply (1)
that problem is the well-studied problem of Padé with the complex variable s1 to f to get the one-
approximation. dimensional Laplace transform
 
Summary fˆs1 t2  ≡ e−s1 t1 f t1 t2  dt1 (22)
0
We regard the discussion so far in this section as pro-
viding support for considering the unified framework which we assume is well defined for Res1  > 0. Then
in (2). Seeing those derivations helps us understand consider the complex variable s1 a constant, and apply
why the framework (2) is natural. Zakian (1969, 1970, (1) a second time with the complex variable s2 to get
1973) and Wellekens (1970) went further, exploiting the two-dimensional Laplace transform
Padé approximation, to obtain specific algorithms,  
considering them as “the solution.” In contrast, we f˜s1 s2  ≡ e−s2 t2 fˆs1 t2  dt2 (23)
0
suggest considering alternative approximants within
this framework. In Sections 4, 5, and 6 we discuss which we assume is well defined for Res1  > 0 and
three very different ways to arrive at approximants, Res2  > 0.
using well-established inversion algorithms. In that Likewise, two-dimensional transform inversion can
way, we demonstrate that the framework may be use- be considered a two-step process, applying the one-
ful without necessarily using Padé approximation. dimensional inversion formula (2) twice: For specified
argument pair t1 t2 , first invert f˜s1 s2  with respect
Framework for Optimization to s2 (inner loop), regarding s1 as constant, to get an
As mentioned earlier, the flexible framework presents approximation for fˆs1 t2 . Then invert the approx-
the opportunity to perform optimization to select the imation for fˆs1 t2  with respect to s1 (outer loop),
“best” procedure, by some criterion. Indeed, one such regarding t2 as constant, getting the desired approxi-
optimization is the Padé approximation performed by mation for f t1 t2 .
Zakian (1970). As an example of a different approach, It is straightforward to extend the unified frame-
we could consider a given value of n and choose work to two dimensions: Following the two-step pro-
the resulting 2n + 2 parameters by doing optimal fit- cedure just described, we have
ting, using families of test functions and mathemat-  
ical programming. To be concrete, suppose that we 1  n2 k2
ˆ ˆ
f s1 t2  ≈ fn2 s1 t2  ≡  ˜
 f s1 (24)
consider m1 test functions fi and m2 arguments of t2 k =0 k2 t2
2
Abate and Whitt: A Unified Framework for Numerically Inverting Laplace Transforms
414 INFORMS Journal on Computing 18(4), pp. 408–421, © 2006 INFORMS

and algorithms , , and  , we find that c = 1 often suf-


n1   fices; see Section 8.
1  k1
f t1 t2  ≈ fn1 t1 t2  ≡ k1 fˆn2 t  (25)
t1 k =0 t1 2
1 4. The Gaver-Stehfest Algorithm
Combining (24) and (27), we obtain the general form Now we start specifying specific one-dimensional
for the inversion inversion algorithms in the unified framework (2). We
start with the Gaver-Stehfest procedure, because it
f t1 t2  ≈ fn1 n2 t1 t2  directly appears in the framework. The Gaver-Stehfest
  algorithm is distinguished by the fact that it does not
1  n1
n2
k1 k2
≡ k1 k2 f˜ (26) use complex numbers; the weights and nodes are real
t1 t2 k =0 k =0
t1 t2 numbers.
1 2

as in (5). The Gaver-Stehfest procedure is based on the se-


We can justify the two-dimensional inversion for- quence of Gaver approximants, fn t% n ≥ 1, derived
mula the same way we did for the one-dimensional by Gaver (1966), which can be written as
  n    
case in Section 2. When we do, we see that the two- n ln2 2n  k n ˆ n + k ln2
dimensional objects are just the product of the two fn t ≡ −1 f 
t n k=0 k t
corresponding one-dimensional objects. For exam- (28)
ple, the two-dimensional scaled delta function is just There is a simple probabilistic derivation, which is
the product of the two one-dimensional delta func- reviewed in Section 8 of Abate and Whitt (1992). Note
tions. Moreover, the two-dimensional linear combi- that the Gaver approximants are directly in the frame-
nation of exponentials is the product of the two work (2).
one-dimensional linear combinations of exponentials. The Gaver approximants can be computed by a
Thus the Laplace transforms are the products of the recursive algorithm, namely,
two one-dimensional Laplace transforms. Hence the k 'k ˆ
one-dimensional Padé approximation can be applied G0 = t
f k'/t 1 ≤ k ≤ 2n
to each dimension in the two-dimensional case. k
Gj
k+1
= 1 + k/jGkj−1 − k/jGj−1
As a consequence of the general two-dimensional
framework in (26), we can apply the three one- 1 ≤ j ≤ n j ≤ k ≤ 2n − j
dimensional routines in Sections 4, 5, and 6 in any n
combination to obtain nine different candidate two- fn t = Gn
(29)
dimensional routines. Even though the Gaver-Stehfest
where ' ≡ ln2.
routine works only with real weights and nodes, we
Unfortunately, however, the convergence fn t →
can apply the Gaver-Stehfest routine in the inner
f t as n →  is slow (logarithmic), so acceleration is
loop with any of these other procedures, because
needed. Accordingly, Stehfest (1970) proposed the lin-
the Gaver-Stehfest routine extends to complex-valued
ear Salzer acceleration scheme. Since the acceleration
functions via (17) and (18).
method is linear, the Gaver-Stehfest algorithm fits in
From the experience in Choudhury et al. (1997), we
the framework (2), just like the Gaver approximants
know that it may be necessary to do extra work for
in (28).
the inner loop to ensure that the intermediate function
This linear acceleration method is reviewed by
fˆs1 t2  has sufficient accuracy to perform the calcula- Valko and Abate (2004), where the Salzer method is
tion in the outer loop accurately. Thus, we anticipate compared to popular nonlinear sequence accelerators.
that we may need n2 > n1 . However, for ease of use, Valko and Abate found that the Salzer scheme per-
it is desirable to have only one parameter n, as in the formed remarkably well, being outperformed only by
one-dimensional case. We can achieve that by letting the Wynn rho algorithm. As an aside, we point out
n1 = n and n2 = cn (27) that the framework (2) is distinguished by encom-
passing linear acceleration techniques, but not nonlin-
for a fixed parameter c, which depends on the spe- ear acceleration techniques, when acceleration is used.
cific routines used in the inner and outer loops. That Here we focus on the original Salzer scheme pro-
is assuming that the system precision required in the posed by Stehfest, because it performs well and
outer loop and the accuracy produced in the inner because it is of the right (linear) form. For any t > 0
loop both grow proportionally to n, which experience and positive integer M, Salzer summation yields the
shows to be approximately so with the algorithms Gaver-Stehfest inversion formula
considered here.  M  
M
n+M n M
What is surprising is that it often suffices to have fg t M = −1 fn t (30)
M! n
c = 1. For the specific two-dimensional algorithms n=1

we consider, built from the three one-dimensional where fn t is given in (28).
Abate and Whitt: A Unified Framework for Numerically Inverting Laplace Transforms
INFORMS Journal on Computing 18(4), pp. 408–421, © 2006 INFORMS 415

Putting (28) into (30) and rearranging the double Thus the efficiency of the Gaver-Stehfest algorithm,
summation, we get measured by the ratio of the significant digits pro-
  duced to the precision required, is
ln2 
2M
ˆ k ln2
fg t M = *f (31) significant digits produced
t k=1 k t eff  ≡
precision required
where
    090M

k∧M
j M+1 M 2j j ≈ ≈ 04 (35)
*k = −1 M+k
(32) 22M
j=k+1/2
M! j j k−j
Let x be the least integer greater than or equal
with x being the greatest integer less than or equal to x.
to x and k ∧ M ≡ mink M.
Inversion formula (31) is in the form (2) with n = Gaver-Stehfest Algorithm Summary
2M, k = k ln2, and k = ln2*k for *k in (32). We If j significant digits are desired, then let M be the
shift the notation from n to M primarily because we positive integer 11j. Given M, set the system pre-
would have to require that n be an even integer. The cision at 22M. Given M and the system precision,
parameter M is natural too, because the algorithm calculate the weights *k , 1 ≤ k ≤ 2M, using (32). Then,
uses M Gaver functionals; the extra M terms in (31) given the transform fˆ and the argument t, calculate
are used to perform the Salzer summation. We consis- the sum fg t M in (31). 
tently use the parameter M for the rest of this paper.
The weights *k in (32) (and thus k in (2)) depend
on both n and k. Since the nodes k = k ln2 are con- 5. The Euler Algorithm
stant multiples of k, there is even spacing of the nodes The Euler algorithm is an implementation of the
along the real line. We remark that Fourier-series method, using Euler summation to
accelerate convergence of the final infinite series; e.g.,

2M see Abate et al. (1999). Since the Fourier-series method
*k = 0 for all M ≥ 1 (33) can also be derived from the Bromwich inversion
k=0
integral in (8), the derivation of the framework (2)
Because of the binomial coefficients in the weights, from the Bromwich inversion integral in Section 2
the Gaver-Stehfest algorithm tends to require high shows that the Fourier-series method fits directly in
system precision in order to yield good accuracy in the framework (2). Since Euler summation is a linear
the calculations. From Abate and Valko (2004), we acceleration algorithm, the full algorithm Euler also
conclude that the required system precision is about fits in the framework (2).
22M when the parameter is M. The precision require- We can also start from the final Euler algorithm
ment is driven by the coefficients *k in (32). Such a itself: we obtain the desired Euler algorithm in the
high level of precision is not required for the compu- framework (2) by fixing the parameters in the Euler
tation of the transform fˆs. algorithm in our previous papers. Referring to the
Abate and Valko (2004) investigated, experimen- algorithm summary on p. 272 of Abate et al. (1999),
tally, the precision produced as a function of the we let old parameter vector l m n A be assigned
parameter M. They found that the answer depends on values 1 M M 2 ln10M/3, where M is a posi-
the transform. From extensive experimentation, Abate tive integer. Here the original l (lower case L) has
and Valko (2004) conclude that about 090M signif- been replaced by 1 (one). We omit that roundoff-
icant digits are produced for f t with good trans- error-control parameter, because we are thinking of
forms. By 090M significant digits, we mean that employing multi-precision software, as in Abate and
Valko (2004). The value of A is taken from equation
f t − fg t M (40) on p. 271 of Abate et al. (1999).

relative error = ≈ 10−090M  (34)
f t With this parameter assignment, the Euler inver-
sion formula to calculate f t numerically for real-
In other words, we are considering the relative error. valued f is
We also need to explain what we mean by good trans-
  
forms. Transforms are said to be “good” (of their 10M/3 
2M
-
class F) if the transforms have all their singularities on fe t M = ,k Re fˆ k (36)
t k=0 t
the negative real axis and the functions f are infinitely
differentiable for all t > 0. If the tranforms are not where
good, then the number of significant digits may not M ln10
-k = + ik ,k ≡ −1k .k (37)
be so great and may not be proportional to M. 3
Abate and Whitt: A Unified Framework for Numerically Inverting Laplace Transforms
416 INFORMS Journal on Computing 18(4), pp. 408–421, © 2006 INFORMS


with i = −1 and integral. We obtain the desired version of Talbot’s
method in the framework (2) by fixing parameters in
1 1 the version of Talbot’s method in Section 3 of Abate
.0 = .k = 1 1 ≤ k ≤ M .2M =
2 2M and Valko (2004). It should be noted that the version
  (38)
−M M of Talbot’s algorithm there is a considerable simplifi-
.2M−k = .2M−k+1 + 2 0 < k < M
k cation of previous versions in the literature.
The Talbot inversion formula to calculate f t
Inversion formula (36) is in the form (2) with n = numerically for real-valued f is
2M, k = 10M/3 ,k and k = -k for ,k and -k in (37) and
(38). Again we let the single parameter be M instead   

2 M−1 !k
fb t M = Re 'k fˆ (42)
of n. As with the Gaver-Stehfest algorithm, about half 5t k=0 t
of the 2M + 1 terms appearing in the sum in (36) are
to accelerate convergence using Euler summation. (subscript b for TalBot), where
Like the Gaver-Stehfest algorithm, there is even
spacing of the nodes -k , but now they are evenly 2M 2k
!0 = !k = cot k/M + i 0 < k < M
spaced on the vertical line s = M ln10/3t instead of 5 5
on the real axis. As with the Gaver-Stehfest algorithm, 1
the weights are real with '0 = e!0 'k = 1 + ik/M1 + cot k/M/2 
2

n − i cot k/M/e!k 0 < k < M
,k = 0 for all even n ≥ 2 (39) (43)

k=0
where again i ≡ −1.
However, now the nodes are complex, so when f is The parameters in the framework (2) are n = M,
real valued, we work with k = !k , and k ≡ 2/5'k for !k and 'k in (43). For
      this fixed-Talbot algorithm, both the weights and the
- - nodes are complex. Now there is uneven spacing of
Re ,k fˆ k = ,k Re fˆ k  (40)
t t the nodes. Now the weights do not sum to zero, but
the sum is small. In particular,
The Euler algorithm tends to be more efficient than
the Gaver-Stehfest algorithm. Given M, the required M−1 

system precision is only about M, but it produces Re 'k ≈ 10−06M  (44)
about 06M significant digits for good transforms. k=0

Thus the efficiency of the Euler algorithm, again mea- As discussed in Abate and Valko (2004), for the
sured by the ratio of the significant digits produced implementation of the Talbot algorithm there may be
to the precision required, is complications with a few complex-function routines.
significant digits produced Those difficulties involve calculating the required
eff  ≡ values of the Laplace transform. Without those dif-
precision required
ficulties, the Talbot algorithm tends to be about as
060M efficient as the Euler algorithm. When the parameter
≈ ≈ 06 (41)
10M is M, about 06M significant digits are produced for a
which is about 3/2 times that of the Gaver-Stehfest good transform, while about M digits of precision are
algorithm. required. Thus the efficiency of the Talbot algorithm is

significant digits produced


Euler Algorithm Summary eff   ≡
If j significant digits are desired, then let M be the precision required
positive integer 17j. Given M, set the system pre- 060M
≈ ≈ 06 (45)
cision at M. Given M and the system precision, cal- 10M
culate the weights ,k and nodes -k using (37) and
just as for the Euler algorithm.
(38). Then, given the transform fˆ and the argument t,
calculate the sum fe t M in (36).  Talbot Algorithm Summary
If j significant digits are desired, then let M be the
6. The Talbot Algorithm positive integer 17j. Given M, let the system pre-
The Talbot algorithm also starts from the Bromwich cision also be M. Given M and the system preci-
integral (8), so it too is destined to fit in the frame- sion, calculate the weights 'k and nodes !k using (43).
work (2). The Talbot algorithm is based on cleverly Then, given the transform fˆ and the argument t, cal-
deforming the contour in the Bromwich inversion culate the sum fb t M in (42). 
Abate and Whitt: A Unified Framework for Numerically Inverting Laplace Transforms
INFORMS Journal on Computing 18(4), pp. 408–421, © 2006 INFORMS 417

7. Performance of the Table 2 A Comparison of the Performance of the Three One-Dimen-


sional Inversion Routines as a Function of M for the Transform
One-Dimensional Algorithms in (46) with Inverse Function in (47)
In this section we briefly discuss the performance of
the three one-dimensional algorithms. We have exten- Significant digits Execution time
sive experience with the Euler algorithm based on M      
our previous work. For the Gaver and Talbot algo-
20 18 13 12 2 4 2
rithms, we draw extensively upon Abate and Valko
30 27 19 18 5 7 4
(2004); see Section 4 for supporting details. The story 50 45 30 30 15 17 10
is quite complicated, so that reference to Abate and 100 91 59 60 90 68 47
Valko (2004) is important for serious study.
Note. The execution time is in milliseconds on a 0.2 GHz CPU.
We first summarize the analysis of efficiency of
the three one-dimensional inversion routines spec-
ified above. Unfortunately, there is no systematic independent of t. However, the results in Table 2 are
error analysis. Moreover, the observed performance quite complicated. In particular, they are nonlinear.
depends on the transform. We describe the perfor- For example, as M goes from 50 to 100,  takes six
mance for good transforms, i.e., those with all singu- times as long, whereas  takes four times as long. This
larities falling on the negative real axis and for which has to do with computing costs at different levels of
the function f is infinitely differentiable (class F in precision, even though the same number of terms are
Abate and Valko 2004). For good transforms, the error being computed. Now compare  and  at the same
is almost independent of t and the transform. For value of M: they both have the same number of terms,
other transforms, that is not true. so we might expect them to take the same time; how-
Twelve examples of transforms in F are listed in ever,  has a complex multiplication for each term,
Table 1 of Abate and Valko (2004). A summary of the whereas  is using only real multiplication.
algorithm efficiency described in the last three sec- Only a few significant digits are required for most
tions appears in Table 1. From our experience, and engineering applications. The very high accuracy we
from Table 1, we conclude that the Euler and Talbot are able to obtain with high system precision for
algorithms are more efficient than the Gaver-Stehfest this example, shown in Table 2, may be impor-
algorithm. However, the Gaver-Stehfest algorithm has tant when we consider high-dimensional multidimen-
the advantage that only real numbers are used. sional inversions, because then we may need greater
We next compare the three algorithms for a specific accuracy in the inner loops in order to obtain required
example, namely the transform precision in the outer loops.
1 There may not need to be great concern about
fˆs = √ (46) the computational effort required to compute the
s+s
weights and nodes, because for multiple applications
which has known inverse the weights and nodes can be computed in advance
√ and stored for easy access (even for multiple values
f t = et erfc t t ≥ 0 (47) of M), but they might also be computed on the fly.
Hence we have investigated the computational effort
where erfc is the complementary error function. The required to compute the weights and nodes for the
performance of the three algorithms as a function of different algorithms. Table 3 shows the time in mil-
the single parameter M is summarized in Table 2. The liseconds (with a 0.2 GHz CPU) required to compute
calculations were performed using the multi-precision the weights and nodes for the three methods , ,
language UBASIC on a 0.2 GHz CPU. and  specified above.
Since the transform in (46) is a good transform (in Since the new version of the Talbot algorithm is rel-
class F, Abate and Valko 2004), the error tends to be atively less tested, we present an additional example

Table 1 A Rough Analysis of Cost and Efficiency for the Three Specified
One-Dimensional Inversion Routines Table 3 A Comparison of the Execution Times for the Weights
and Nodes (in Milliseconds with a 0.2 GHz CPU) for the
Significant Three Specified One-Dimensional Inversion Routines
No. of System precision digits produced
Inversion transform required for good Efficiency M Euler Talbot Gaver-Stehfest
routine evaluations decimal digits transform (SD/prec.)
25 ≈0 6 12
Gaver-Stehfest 2M 22M 090M 0.4 50 1 25 100
Euler 2M + 1 10M 060M 0.6 100 4 160 1200
Talbot M 10M 060M 0.6 200 20 1600 19000
Abate and Whitt: A Unified Framework for Numerically Inverting Laplace Transforms
418 INFORMS Journal on Computing 18(4), pp. 408–421, © 2006 INFORMS

Table 4 Significant Digits Produced by the One-Dimen- 8. Candidate Two-Dimensional


sional Talbot Algorithm as a Function of t and
M for the Example in (48) and (49) Algorithms
In this section we combine the three one-dimensional
M inversion algorithms , , and  specified in Sections
t 10 20 40 100 200 4, 5, and 6 to construct nine two-dimensional inver-
sion algorithms. In each case, the two-dimensional
10−8 1 10 23 59 119 inversion routine can be expressed in the common
10−6 6 12 23 59 119
form (26), but in some cases further simplification can
10−2 6 12 23 59 119
10−1 6 12 23 59 119 be obtained by exploiting properties of complex con-
100 6 11 23 59 119 jugates. We display the possible algorithms below in
101 5 11 22 58 118 Table 5.
102 5 10 21 57 118 In Table 5 we use the notation in Sections 4–6. In
104 3 9 20 55 114 particular, we use the single parameter M instead of
106 2 8 19 54 113
108 1 7 18 53 112
n. The outer loop has parameter M and the inner loop
has parameter cM, so there remains only the single
parameter M.
examining its performance. We apply the Talbot algo- We also use the special notation for the nodes and
rithm to the transform weights introduced in Sections 4–6. Thus, *k are the
1 Gaver weights in (32), ,k and -k are the Euler weights
fˆs = √ √ (48) and nodes in (37) and (38), and 'k and !k are the Tal-
s+ s+1
bot weights and nodes in (43).
which has known inverse It is understood that the weights and nodes in the
1 − e−t outer loop depend on the parameter M, while the
f t = √ t ≥ 0 (49) weights and nodes in the inner loop depend on
4t 3 the parameter cM, where c is specified at the left,
Numerical results for a wide range of t and M are based on our experience with a collection of good
given in Table 4. Again we see excellent performance. transforms (the class F from Abate and Valko 2004).

Table 5 The Nine Two-Dimensional Inversion Routines Based on the Three One-Dimensional Routines:
Gaver-Stehfest  , Euler  , and Talbot 
  

2 ln2 M−1 
2M k k ln2
 M M: Re
k1 k2 f˜ 1  2
5t1 t2 k =0 k =1
t1 t2
1 2
 M−1     
2  M−1  k k k ¯k
 M M: Re
k1
k2 f˜ 1  2 +
k2 f˜ 1  2 ,
25t1 t2 k =0 k =0
t1 t2 t1 t2
1 2
  
10M/3 ln2 2M 
2M k1 k2 ln2
 M M: k1 k2 Re f˜ 
t1 t2 k =0 k =1
t1 t2
1 2
    
10 M/3  2M  M−1 k1 k2 k1 ¯k2
 M M: k1 Re
k2 f˜  +
k2 f˜ 
5t1 t2 k =0 k =0
t1 t2 t1 t2
1 2
     
10 M/3 M−1  
2M k k k ¯k
 M M: Re
k1 k2 f˜ 1  2 + f˜ 1  2
5t1 t2 k1 =0 k =0
t1 t2 t1 t2
2
   
2 ln2 2M 
3M−1
k ln2 k2
 M 3M: k1 Re
k f˜ 1 
5t1 t2 k =1 k =0 2 t1 t2
1 2
 
ln22 
2M 
4M
k ln2 k2 ln2
 M 2M: k1 k f˜ 1 
t1 t2 k =1 k =1 2 t1 t2
1 2
    
10 2M/3  
2M 2M k1 k2 k1 ¯k2
 M M: k1 k2 Re f˜  + f˜ 
2t1 t2 k =0 k =0
t1 t2 t1 t2
1 2
   
10 ln2 
M2M 
6M
k ln2 k2
 M 3M: k1 k Re f˜ 1 
t1 t2 k =1 k =0 2 t1 t2
1 2

Notes. The value of c has been specified based on experience with good transforms. The routines are ordered
according to the number of transform evaluations required. A prime appears on the nodes and weights in the inner
loop in the three cases in which c = 1, indicating that these parameters depend on cM, not M.
Abate and Whitt: A Unified Framework for Numerically Inverting Laplace Transforms
INFORMS Journal on Computing 18(4), pp. 408–421, © 2006 INFORMS 419

When c = 1, which occurs in only three cases, the cor- Table 7 A Comparison of the Performance of the Nine Two-Dimen-
responding parameters in the inner loop are desig- sional Inversion Routines as a Function of M for the Transform
nated by the addition of a  (prime). For example, with in (50) with Inverse Function in (51)
the all-Gaver routine M2M, *k depends on M, Significant digits Execution time
while *k depends on 2M. Similarly, in M3M, *k M M
again depends on M, but ,k and -k depend on 3M.
Routine 10 20 30 50 10 20 30 50
However, much to our surprise, in six of the nine two-
dimensional algorithms we found that it suffices to  M M 6 12 18 30 0.05 0.20 0.6 3
let c = 1, making the number of computations in the  M M 6 12 18 30 0.04 0.15 0.4 2
 M M 6 12 17 27 0.05 0.41 1.3 7
inner loop no greater than in the outer loop.
 M M 7 13 19 30 0.07 0.29 0.8 4
The four algorithms combining  and  exploit  M M 7 13 19 30 0.07 0.28 0.8 4
relations involving complex conjugates, so there is  M 3M 8 16 24 40 0.13 1.00 3.6 21
an extra term in each sum, with the entire quantity  M 2M 9 13 17 28 0.19 0.80 2.8 19
divided by 2.  M M 6 14 18 30 0.14 0.71 2.0 9
 M 3M 8 16 24 39 0.27 1.90 7.0 40

9. Performance of the Note. The execution time is in seconds on an 0.2 GHz CPU.

Two-Dimensional Algorithms
In this section we describe the performance of with inverse function
the two-dimensional inversion routines. Paralleling  1/2
2
Table 1, in Table 6 we analyze the efficiency of the f t1 t2  = √ t12 + t22 − t2  (51)
nine routines specified in Section 8. We order the algo- 
rithms according to the approximate number of trans- Numerical results for this first example are given in
form evaluations required, because that seems to be Table 7. We consider all nine methods in Table 5.
the decisive factor. The algorithms at the top, requir- The execution time reflects the number of transform
ing fewer transform evaluations, tend to be most effi- evaluations, but it also reflects the required precision
cient.
setting. For example,   and   require about the
Now we consider two concrete examples in detail,
same number of transform evaluations, 2M 2 , but the
both taken from Ditken and Prudnikov (1962). Both
execution time for   is higher because the precision
are good transforms (in the class F in Abate and Valko
setting is higher. The precision setting is 22M for  ,
2004). (We have examined about ten examples from
but only M for   . In contrast,  and   have both
Ditken and Prudnikov (1962); these two are repre-
the same number of evaluations, 4M 2 , and the same
sentative. We also have conducted experiments on
precision setting, so that the overall execution time
queueing transforms. Applications to queues will be
is consistently about the same. On the other hand,
discussed in Abate and Whitt 2005.) The first example
here is the two-dimensional Laplace transform   and  have the same precision setting, but  
  evaluates half as many terms as  ; accordingly, the
ˆ 1 s1 execution time for   is about half that for  .
f s1 s2  = √ 1−
(50)
s1 s2 s1 s1 + s2 + 2s1 s2 Our second example is the two-dimensional Laplace
transform
Table 6 A Rough Analysis of Cost and Efficiency for the Nine Two-


exp 1/ s s + 1
fˆs1 s2  =
2 1
Dimensional Inversion Routines
(52)
s2 s 1 + 1
Inversion Approx. no. of Sig. digits produced Efficiency
routine transform evaluations for good transform (sd’s/evals.) with inverse function
 M M 2M 2 06M 0300/M
e−t1 
1/2
 M M ∗2M 2 06M 0300/M f t1 t2  = √ I 8 t1 t2 (53)
 M M 4M 2 06M 0150/M
t1 0
 M M ∗4M 2 06M 0150/M
 M M ∗4M 2 06M 0150/M
where I0 is the modified Bessel function; see p. 66 of
Magnus et al. (1966). Numerical results for this second
 M 3M 6M 2 08M 0133/M
 M 2M 8M 2 06M 0075/M
example are given in Table 8.
 M M ∗8M 2 06M 0075/M From Tables 7 and 8, we conclude that the algo-
 M 3M 12M 2 08M 0067/M rithms in the upper part of the tables, requiring fewer
Note. The routines have been ordered according to the approximate number
transform evaluations, are most efficient. Rough gen-
of required transform evaluations. The asterisks indicate values that have eralizations would be: (1) it is not so good to use the
been multiplied by two because there is a second evaluation involving the Gaver-Stehfest routine in the outer loop and (2) the
complex conjugate. Talbot routine tends to be efficient wherever it is used.
Abate and Whitt: A Unified Framework for Numerically Inverting Laplace Transforms
420 INFORMS Journal on Computing 18(4), pp. 408–421, © 2006 INFORMS

Table 8 A Comparison of the Performance of Nine Two-Dimensional In Sections 7 and 9 we reported results on the
Inversion Routines as a Function of M for the Transform in performance of the algorithms. In Section 7 we con-
(52) with Inverse Function in (53)
cluded that the Euler and Talbot one-dimensional
Significant digits Execution time algorithms are about equally efficient, as measured by
M M the ratio of the number of significant digits produced
Routine 10 20 30 50 10 20 30 50
to the number of digits of machine precision required,
while these two are about 15 times as efficient as
 M M 7 13 19 31 0.08 0.31 11 7 the Gaver-Stehfest one-dimensional algorithm. How-
 M M 6 12 18 30 0.05 0.19 06 3
ever, the Gaver-Stehfest algorithm has the advantage
 M M 6 13 19 28 0.08 0.63 22 13
 M M 7 12 18 30 0.10 0.37 11 5 of working with only real numbers.
 M M 7 13 19 30 0.09 0.35 11 5 Investigations of the two-dimensional algorithms in
 M 3M 9 18 28 46 0.22 1.86 72 48 Section 9 indicate that all nine combinations of the
 M 2M 9 13 17 26 0.27 2.04 84 72 three one-dimensional routines can be effective two-
 M M 6 13 18 30 0.24 1.12 36 17
dimensional algorithms. However, we encountered
 M 3M 9 17 22 37 0.35 2.73 112 81
difficulties calculating values of two-dimensional
Note. The execution time is in seconds on a 0.2 GHz CPU. transforms when the Talbot method  is used in
the inner loop, because of problems evaluating trans-
However, there is a difficulty with the Talbot algo- forms for conjugate s. Initial results indicate that the
rithm, which does not have to do with efficiency. It best two-dimensional algorithms are the combina-
is caused by the fact that the Talbot routine involves tions  ,  and  . Evidently it can be advanta-
both complex weights and complex nodes. We found geous to use different one-dimensional routines in the
that we sometimes encountered difficulties evaluat- inner and outer loops.
ing transform values when the Talbot algorithm was Much remains to be done. First, as is often the
used in the inner loop. That is because of difficulties case with numerical transform inversion (and many
in the software (e.g., Mathematica) properly evaluating other numerical methods), a systematic error anal-
complex-valued functions of two complex variables, ysis is lacking. The bounds for the Fourier-series
e.g., involving the square-root function. method with Euler summation in Abate and Whitt
Thus, at this time, we would not recommend using (1995), O’Cinneide (1997), and Sakurai (2004) are the
the Talbot algorithm in the inner loop, or at least exception. There is a need for much more testing. It
being aware that there might be problems involv- remains to understand better how the algorithms per-
ing computation with complex numbers and func- form on different classes of transforms. More work is
tions. That leaves  , , and   as the top-ranked needed, extending Abate and Valko (2004), to under-
two-dimensional routines. None of these uses the same stand how the algorithm performance depends on the
one-dimensional routine in both loops. However, all nine structure of the transform. It remains to see what com-
methods are effective. Overall, our study indicates the binations of the three one-dimensional routines will
unified framework has promise. be especially effective for high-dimensional inversion
problems.
10. Conclusions It also remains to consider other kinds of trans-
In this paper we have proposed the general unified forms. We have found that essentially the same uni-
framework in (2) for numerically inverting Laplace fied framework applies to generating functions; we
transforms. We pointed out that the flexible frame- plan to elaborate on that elsewhere. Thus, just as in
work makes it possible to perform optimization in Choudhury et al. (1994a), the unified framework will
order to select specific inversion routines for special apply to multidimensional transforms, where each
classes of functions, but it remains to pursue that dimension may involve either a generating function
approach. In Sections 4, 5, and 6 we showed that the or a Laplace transform.
Gaver-Stehfest method, , the Fourier-series method
with Euler summation, , and a version of Talbot’s
method,  , all fit into this common framework. References
In Section 3 we observed that the unified frame- Abate, J., P. P. Valko. 2004. Multi-precision Laplace inversion. Inter-
work extends directly to yield a corresponding frame- nat. J. Numer. Meth. Engrg. 60 979–993.
work for numerically inverting multidimensional Abate, J., W. Whitt. 1992. The Fourier-series method for inverting
transforms of probability distributions. Queueing Systems 10
Laplace transforms. In Section 8 we observed that
5–88.
the three specific one-dimensional algorithms we Abate, J., W. Whitt. 1995. Numerical inversion of Laplace trans-
introduced in the framework (2) produce nine can- forms of probability distributions. ORSA J. Comput. 7 36–43.
didate two-dimensional algorithms, all in the two- Abate, J., W. Whitt. 1997. Asymptotics for M/G/1 low-priority
dimensional framework (5). waiting-time tail probabilities. Queueing Systems 25 173–223.
Abate and Whitt: A Unified Framework for Numerically Inverting Laplace Transforms
INFORMS Journal on Computing 18(4), pp. 408–421, © 2006 INFORMS 421

Abate, J., G. L. Choudhury, W. Whitt. 1998. Numerical inversion of Evans, G. A., K. C. Chung. 2000. Laplace transform inversion using
multidimensional Laplace transforms by the Laguerre method. optimal contours in the complex plane. Internat. J. Comput.
Performance Eval. 31 229–243. Math. 73 531–543.
Abate, J., G. L. Choudhury, W. Whitt. 1999. An introduction to Feller, W. 1971. An Introduction to Probability Theory and Its Applica-
numerical inversion and its application to probability mod- tions, 2nd ed. Wiley, New York.
els. W. Grassman, ed. Computational Probability. Kluwer, Boston, Gaver, D. P. 1966. Observing stochastic processes and approximate
MA, 257–323. transform inversion. Oper. Res. 14 444–459.
Audis, E., W. Whitt. 2006. Power algorithms for inverting Laplace Graf, U. 2004. Applied Laplace Transforms and z-Transforms for Scien-
transforms. Columbia University, New York. Available at tists and Engineers: A Computational Approach Using a Mathemat-
http://www.columbia.edu/~ ww2040/recent.html. ica Package. Birkhauser-Verlag, Berlin, Germany.
Baker, G. A., P. Graves-Morris. 1996. Padé Approximants, 2nd ed.
Hwang, C., M. J. Lu. 1999. Numerical inversion of 2-D Laplace
Encyclopedia of Mathematics and Its Applications, Vol. 59.
transforms by fast Hartley transform computations. J. Franklin
Cambridge University Press, Cambridge, UK.
Inst. 336 955–972.
Bingham, N. H., C. M. Goldie, J. L. Teugels. 1987. Regular Vari-
Kao, E. P. C. 1997. An Introduction to Stochastic Processes. Duxbury
ation. Encyclopedia of Mathematics and Its Applications, Vol. 27.
Press, New York.
Cambridge University Press, Cambridge, UK.
Choudhury, G. L., W. Whitt. 1997. Probabilistic scaling for the Magnus, W., F. Oberhettinger, R. P. Soni. 1966. Formulas and Theo-
numerical inversion of nonprobability transforms. INFORMS rems for the Special Functions of Mathematical Physics. Springer,
J. Comput. 9 175–184. New York.
Choudhury, G. L., K. K. Leung, W. Whitt. 1995a. Calculat- Murli, A., M. Rizzardi. 1990. Algorithm 682. Talbot’s method for
ing normalization constants of closed queueing networks by the Laplace inversion problem. ACM Trans. Math. Software 16
numerically inverting their generating functions. J. ACM 42 158–168.
935–970. Narayanan, G. V., D. E. Beskos. 1982. Numerical operational meth-
Choudhury, G. L., K. K. Leung, W. Whitt. 1995b. An inversion ods for time-dependent linear problems. Internat. J. Numer.
algorithm to compute blocking probabilities in loss networks Methods Engrg. 18 1829–1854.
with state-dependent rates. IEEE/ACM Trans. Networking 3 O’Cinneide, C. A. 1997. Euler summation for Fourier series and
585–601. Laplace transform inversion. Stochastic Models 13 315–337.
Choudhury, G. L., D. M. Lucantoni, W. Whitt. 1994a. Multidimen- Piessens, R. 1969. New quadrature formulas for the numerical
sional transform inversion with applications to the transient inversion of the Laplace transform. BIT 9 351–361.
M/G/1 queue. Ann. Appl. Probab. 4 719–740. Piessens, R. 1971. Gaussian quadrature formulas for the numerical
Choudhury, G. L., D. M. Lucantoni, W. Whitt. 1994b. The transient inversion of the Laplace transform. J. Engrg. Math. 5 1–9.
BMAP/G/1 queue. Stochastic Models 10 145–182. Sakurai, T. 2004. Numerical inversion of the Laplace transform of
Choudhury, G. L., D. M. Lucantoni, W. Whitt. 1997. Numerical solu- functions with discontinuities. Adv. Appl. Probab. 36 616–642.
tion of Mt /Gt /1 queues. Oper. Res. 45 451–463.
Stehfest, H. 1970. Algorithm 368: Numerical inversion of Laplace
Davies, B. 2002. Integral Transforms and Their Applications, 3rd ed. transforms. Comm. ACM 13 47–49, 624.
Springer, New York.
Talbot, A. 1979. The accurate inversion of Laplace transforms.
Davies, B., B. Martin. 1979. Numerical inversion of the Laplace J. Inst. Math. Appl. 23 97–120.
transform: A survey and comparison of methods. J. Computa-
tional Phys. 33 1–32. Valko, P. P., J. Abate. 2004. Comparison of sequence accelerators for
the Gaver method of numerical Laplace transform inversion.
Davis, P. J., P. Rabinowitz. 1984. Methods of Numerical Integration. Comput. Math. Appl. 48 629–636.
Academic Press, New York.
Valko, P. P., B. L. Vojta. 2001. The List. Bibliography of papers
Ditken, V. A., A. P. Prudnikov. 1962. Operational Calculus in Two
on numerical transform inversion and its application. http://
Variables and Applications. Pergamon, New York. (English trans-
pumpjack.tamu.edu/∼valko.
lation of 1958 Russian edition.)
Vlach, J. 1969. Numerical method for transient responses. J. Franklin
Doetsch, G. 1974. Introduction to the Theory and Application of the
Inst. 288 91–113.
Laplace Transformation. Springer, New York.
Dubner, H., J. Abate. 1968. Numerical inversion of Laplace trans- Wellekens, C. J. 1970. Generalization of Vlach’s method for the
forms by relating them to the finite Fourier cosine transform. numerical inversion of the Laplace transform. Electronic Lett. 6
J. ACM 15 115–123. 742–744.
Duffy, D. G. 1993. On the numerical inversion of Laplace trans- Zakian, V. 1969. Numerical inversion of Laplace transform. Elec-
forms: Comparison of three new methods on characteristic tronic Lett. 5 120–121.
problems from applications. ACM Trans. Math. Software 19 Zakian, V. 1970. Optimisation of numerical inversion of Laplace
333–359. transforms. Electronic Lett. 6 677–679.
Evans, G. A. 1993. Numerical inversion of Laplace transforms using Zakian, V. 1973. Properties of IMN approximants. P. R. Graves-
optimal contours in the complex plane. Internat. J. Comput. Morris, ed. Padé Approximants and Their Applications. Academic
Math. 49 93–105. Press, New York, 141–144.

You might also like