Sunteți pe pagina 1din 11

Rheologica Acta

Rheol Acta 32:311-321 (1993)

Generating line spectra from experimental responses.


Part h Relaxation modulus and creep compliance
I. Emri 1 and N.W. Tschoegl 2
University of Ljubljana, Ljubljana, Slovenia
2California Institute of Technology, Pasadena, California, USA

Abstract: We describe a recursive computer algorithm which generates line spectra from relaxation modulus or creep compliance data without producing
negative spectrum lines. We apply the algorithm here to data read from
mathematical models for the relaxation modulus. Since these data were thus free
of the usual experimental error, we could use a relatively simple form of the basic
algorithm that is applicable also to smoothed data. The spectra faithfully
reproduced the input functions and may serve for data storage as well as for
predicting other experimental responses.
Key words."Creep compliance - line spectrum - relaxation modulus - relaxation spectrum - retardation spectrum

Introduction

In the response of a linearly viscoelastic material to


a strain excitation, the relaxation spectrum, H ( z ) ,
contains complete information on the time-dependent
part of the response. In the response to a stress excitation it is the retardation spectrum, L ( z ) , which contains the information. In either case the total response
then consists of the appropriate viscoelastic constants
(such as the equilibrium modulus, or the glass compliance and the steady-flow fluidity), in addition to
the integral over the spectrum multiplied by a kernel
function characteristic of the type of excitation
chosen to elicit the response. I f this is a strain or a
stress as a step function of time, the kernel is the exponential function, exp ( - t/r), and the result of the
integration is the relaxation modulus in the first case,
and the creep compliance in the second case. Thus,
once the spectrum and the viscoelastic constants are
known, it is possible, in theory, to generate the
response to any desired type of excitation.
The spectrum itself is not accessible by direct experiment (Tschoegl, 1989; pp. 157-158). From a
mathematical model of a response curve it can often
be calculated exactly (Tschoegl, 1989; p. 196). However, f r o m experimental data the spectrum is
necessarily obtained as an approximation to the unknown 'true' spectrum whose existence is but assumed. Commonly, one obtains approximations to the

spectra f r o m experimental data by numerical or


graphical differentiation or by the use of finite difference calculus (Tschoegl, 1989; Sect. 4.2 and
Sect. 4.3). These methods are based on the canonical
(i.e., integral) representations (Tschoegl, 1989;
p. 160) of the experimental response and yield sets of
values sampled f r o m the approximations to the continuous spectra.
Another approach is based on the representation of
the experimental response by discrete canonical
models (Tschoegl, 1989; pp. 134, 160) leading to
distributions of line spectra. Once a line spectrum is
available, any response curve within the same
representational group (i,e., within relaxation or
creep behavior) can be generated readily. Line spectra
obtained f r o m time-dependent and frequency-dependent data can easily be combined into a single spectrum covering a wide range of respondance times.
Thus, a line spectrum is also an excellent means for
the storage of experimental data.
Determination of a line spectrum is essentially a
curve fitting procedure. We must distinguish between
two possible fits. The first will produce the best
mathematical fit, but is bound to produce some
negative spectrum lines. Such lines are physically
unacceptable and can create havoc in any interconversion, particularly when converting between relaxation
and retardation behavior. The second fit eliminates
all negative spectrum lines (zero lines are physically

312
acceptable). Such a fit will not produce the best fit to
the datum points, but will more closely approach the
intrinsic material property. No line spectrum - produced by whatever method - is ever the true spectrum. This necessarily remains unknown. Furthermore, no such spectrum is unique because its exact
nature depends on the parameters introduced to obtain it. The success of a spectral distribution is measured by comparing the original experimental data
with those reconstructed from the distribution.
Unfortunately, attempts to develop suitable
methods for obtaining well-determined line spectra
from which the input function can be faithfully
reconstructred, have hitherto not met with unqualified success. Among the better known older
methods of this type (Tschoegl, 1989; Sect. 3.6, p.
136ff) are Procedure X of Tobolsky and Murakami
(1959), the Collocation method of Schapery (1962),
and an extension of it by Cost and Becker (1970) called the Multidata method. The two last named require
matrix inversion. Both are likely to generate negative
modulus or compliance values in portions of the line
spectrum. If positive lines appropriately compensate
such negative lines, the occurrence of the latter may
not unduly affect the reconstruction of the input
curve. However, this compensation depends on the
exact choice of the input parameters. This makes it
virtually impossible to generate any other experimental responses from a "compensated" spectrum with
any degree of confidence.
Due undoubtedly to the easy availability of sufficiently powerful computers, the last several years
have seen an increase in interest in methods for computing spectral distributions. The literature now contains several more recent papers devoted to this subject (Baumgaertel and Winter, 1989; Elster and
Honercamp, 1991, 1992; Elster, Honercamp and
Weese, 1991; Honercamp, 1989; Honercamp and
Weese, 1989; Janzen and Scanlan, 1993; Kamath and
Mackley, 1989; Laun, 1986, Orbey and Dealy, 1991).
We wish to make our own contribution to it by offering a recursive computer algorithm specifically
designed to avoid negative spectrum lines. The method of Baumgaertel and Winter (1989), e.g., also
avoids negative spectrum lines, albeit at the expense
of reducing the density of the lines. However, we
refrain here from a detailed review o f earlier papers
by other authors because an extensive comparison is
currently in preparation (Tschoegl and Emri, to be
submitted). We do point out, however, that our
algorithm does not require matrix inversion and that
- in contrast to all other methods of which we are
aware - it computes each spectrum line utilizing only

Rheologica Acta, Vol. 32, No. 3 (1993)


well-defined subsets of the complete set of experimental data. The choice of these subsets is governed by
the form of the kernel functions in the canonical
representations of the experimental response functions. Consequently, there exist several forms of the
basic algorithm, one for each kernel function. Our
algorithms (we use the singular or the plural as appropriate) thus avoid potential sources of instabilities in
the computations.
In this paper, we introduce the algorithm in the
form in which it is applied to simulated step response
curves, i.e., to continuous mathematical functions
having all the characteristics of the relaxation
modulus or the creep compliance demanded by the
theory of linear viscoelasticity (Tschoegl, 1989; Sect.
6.1, pp. 314-326). A companion paper (Tschoegl
and Emri 1993a) deals with the determination of the
spectra from simulated storage and loss functions. A
third paper (Tschoegl and Emri, 1992) takes up the
problem of interconverting between relaxation and
retardation line spectra. In these three papers, we
apply the algorithm to data sets obtained by sampling
continuous theoretical curves which - being free of
experimental error - are either monotone non-increasing [e.g. G (t), J' (co)], monotone non-decreasing
[e.g. G'(og), J(t)], or both [e.g. G"(co), J"(co)]. This
simplifies the algorithm, the presentation of its basic
nature, and the demonstration of its power. A fourth
paper [Emri and Tschoegl, 1993b) deals with the application to experimental data exhibiting considerable
scatter as is usual, e.g., when the time scale has been
extended through time-temperature superposition. In
that paper, we propose a modification of the
algorithm in which the relative rather than the absolute error forms the basis of the criterion for exiting
from an iteration. It must be pointed out, however,
that the simpler forms of the algorithm that we report
in this and in the companion paper are by no means
restricted to simulated data. They can, e.g., be applied successfully to smoothed data (Emri und
Tschoegl, 1993 b). Smoothing of the data with an efficient smoothing tool such as the cubic spline function
(Emri and Tschoegl, 1993b) is almost universally
beneficial. We have, however, applied the simpler
forms of the algorithm described here with excellent
results also to unsmoothed data that contained
moderate experimental error.
Theoretical

The form of the basic algorithm we discuss here


determines a line spectrum from a step response function such as G(t) or J(t). It obtains the strength of the

Emri and Tschoegl, Generating line spectra from experimental responses


line associated with each of a set of predetermined
respondance times (relaxation or retardation times)
recursively. Although it does not require that the
respondance times be equally spaced, we use equalspacing for convenience.
To understand how the algorithm does its work, let
us first consider the reverse calculation, i.e., the reconstruction of the relaxation modulus from a
discrete distribution of spectrum lines. Let this distribution, for simplicity's sake, consist of four lines,
equally spaced on the logarithmic time scale, and let
gi(t) = giexp ( - t/r,)

(1)

{i = 1,2,3,4}

be the time-dependent contribution to the relaxation


modulus provided by the strength of the ith normalized spectrum line, gi, multiplied by the kernel function, exp ( - t / r , ) . The spectrum is shown in Fig. 1 a.
Figures l b to l e display the four gi(t) =
gi exp ( - t/r,) as functions of log t. Each gi(t) can be
subdivided into three parts: the "glassy" region in
which it is virtually constant with value gi because
here the kernel function approaches unity; the "transition" region in which it is time-dependent, and the
"equilibrium" region in which it vanishes. What now
are the contributions which the gi(t)'s make to the

(normalized) relaxation modulus, g(t)? We consider


g(t) in the vicinity of t = r 2 (see Fig. lc). Let us
define this vicinity for the moment simply as the
region tt<_t<_t~ without specifying tt or tu (the
subscripts refer to the lower and the upper bounds of
the region). Also, for the moment, we assume that
there is only a single spectrum line within this region.
The contribution of gl (t) then essentially vanishes,
while those of g3(t) and g4(t) are just g3 and g4
because the value of the exponentials is essentially
unity. The contribution of g2(t), however, remains
g2 exp ( - t/r2). The value of the (normalized) relaxation modulus in the vicinity of t = r2 is then given by
the sum of all contributions, i.e., by
gut<_ t_< t~) = 0 + g2 exp [ - (tt_< t_< t~)/r2]

(2)

+g3 +g4 .

Figure i f illustrates this. Equation (2) represents the


Wiechert (or generalized Maxwell) model for a
rheodictic 1) material characterized by four relaxation
times. It emphasizes the point that the contributions
of "far" spectrum lines to the right of the vicinity are
actually constant, while the "far" spectrum lines to
the left of the vicinity vanish. Thus, in the general
case,
g

a.

313

(tl-< t_< tu) = 0 + gk exp [ -- (tt <--t <_t u)/r k]

.6

i=N

.2
O

~ 0
.6

(WINOOW lh
i

(3)

~ gi ,
i=k+l

where N is the total number of spectrum lines. For


any datum point within the vicinity of the farthestright spectrum line, gN, (see Fig. i e) we simply have

(WlNOOW llz

~.2

~" A I

g(tt<--t<-tu)=gNexp[--(tl<--t<--tu)/rU]

(4)

(WINDOW lh

~ '2I

Knowing r N and any datum point within the vicinity


of r N, we can obtain gu directly from Eq. (4). This,
therefore, may be taken as the starting relation for a
recursive calculation of the desired discrete distribution of spectrum lines. Equation (3) then furnishes the
recursion equation as

,
(WINDOW lh

__ 68it ...............
02

"~

........... r

,4

i=N

g(tt<-t<-tu) -

.2

-4

-3

-2

-1

Io0 (fir)
Fig. l a. Discrete relaxation spectrum consisting of four
(normalized) spectrum lines; b, c, d, e. gih(t-ri) and
gi exp ( - t/zt) as functions of log t; f. (normalized) relaxation modulus, g(t), as function of log t

gk =

gi

i=k+l

exp [ - (tt_< t < tu)/rk]

(5)

1) The term rheodictic refers to a material showing


steady-state flow. Arrheodictic then denotes a material
which does not [Tschoegl, 1989, p. 93].

314

Rheologica Acta, Vol. 32, No. 3 (1993)

Hence, we derive the strength of each spectrum line


from a single datum point selected at a value of t
restricted to tl<_t<_t u. It remains to show why the
calculation of each line must be restricted to a datum
point within the "vicinity", and what, exactly, we
mean by that hitherto loosely defined term.
To deal with these questions, we replace the kernel
functions, exp (--t/rk), by unit step functions of
time, h (rk--t), which cut off all contributions above
t = rk. The step functions may be considered to be
crude approximations to the exponential functions.
Figures 1 b, i c, 1 d, and 1 e show them as dashed lines
next to the full lines representing the exponential
functions. The step functions exhibit no "transition"
regions. Hence, the (normalized) relaxation modulus
becomes
k-1

g(t)=

~ gih(ri-t)+gkh(rk-t)
i=1
i=N

gi h(T i - t ) .

(6)

i=k+l

In this expression, we have separated out from the


total sum the k t h spectrum line as that line on which
we focus our attention. If now there exists a datum
point at t = rk, then it follows from Eq. (6) that
i=N

gk=g(t)[t=r~ -

gi.

(7)

i=k+l

If, however, we do not find a datum point at t = rk,


then there would be no way to calculate gk. Thus, if
we replace the exponential functions with step functions it becomes clear that we should not use datum
points which lie either to the left or to the right of
t = r k for the evaluation of the k t h spectrum line.
That this is so is seen even more clearly when we examine Fig. I f. If, for example, we use the datum
point at t = rl, denoted by the filled circle, to
calculate g4, we would erroneously obtain

g4=g(t)lt=~l ,

function of time. As we have just shown, the discontinuous nature of the step function at t = Tk allows
the calculation of the kth spectrum line only from the
single datum point on the relaxation curve which lies
at t = Zk (or, at most, from one very close to it). By
contrast, the exponential function replaces the sharp
cut-off at t = rk by a continuous "transition" region
on the logarithmic time scale, This vastly reduces the
error involved in using neighboring datum points in
the calculation and is important because, in any practical application, the location of the "coinciding"
data point is not known. The advantage is valid in the
reverse case also. To model the relaxation modulus
with exponential kernels instead of step functions requires, for the same error, a much smaller number of
elements.
The salient conclusion to be drawn from these
observations is that datum points should be selected
only within a window covering essentially the "transition" region. This window is what we have earlier
called the "vicinity". The exponential kernel function
can model the relaxation modulus effectively only in
the transition region where the first derivative is significantly larger than zero. Figure 2 shows that the exponential kernel function well satisfies this condition
in the region from about log t/rk = - 0.6 to 0.4. This
region is the Boundary Window. Further below, we
introduce a second window, the Modeling Window. It
is this window in which the algorithm models the transient response most efficiently. The span of this window will be "bounded" by the first window which
takes its name from this property. However, to save
space in the figures, we will refer to these windows as
Window 1, and Window 2, respectively.
In this region the exponential function, when plotted semilogarithmically, can be well modeled by a

WINOOW 1

(8)
"-'- .6

'i"

as shown by the partially filled circle and the vertical


line in the figure. Modeling of the relaxation modulus
then would require spectrum lines with negative
values in the time domain between rl and r4 to compensate for the error.
The use of step functions instead of exponential
kernels effectively reduces the vicinity to a single
point. It can now be recognized as an advantage that
the kernel function is an exponential, and not a step

OL

1.5
-.5
log (t/v)
Fig. 2. Definition of Window 1 (the Boundary Window)

-2

-1.5

-!

Emri and Tschoegl, Generating line spectra from experimental responses


straight line. The reason for selecting the particular
limits of - 0 . 6 and 0.4 will become clear a little later.
The window is not equally spaced around log t / r k
because the exponential kernel function is not "selfc o n g r u e n t " (Tschoegl, 1989; p. 318), i.e., the upper
and lower halves of the curve produced by plotting
log g ( t ) vs log ( t / r ) cannot be brought into superposition with each other by reflection and translation. In
the companion paper (Tschoegl and Emri, 1993 a) we
shall prove that this is actually an advantage not
possessed by the self-congruent kernels of the storage
and the loss functions.
Outside of Window 1 the exponential kernel
behaves effectively as a step function. As discussed in
connection with Eqs. (7) and (8), an attempt to
calculate a spectrum line outside the "boundary window" may produce a line that is either larger or
smaller than it should be. The algorithm then tries to
compensate by making the next line smaller or larger.
We can now distribute the complete set of datum
points over several windows, each of which comprises
the location of one of the preselected spectrum lines.
A window of the stated width allows us to calculate
spectrum lines separated by exactly one decade on the
logarithmic time scale. We have also found it convenient when running the algorithm in double precision
on a computer. To cover the entire range of datum
points on the logarithmic time scale, one needs to
preselect at least one spectrum line per window (i.e.,
per decade of time). Each window must contain at
least one datum point. However, the datum points
need not be spaced equally as long as each is in a
separate window. If enough datum points are available, the number of spectrum lines per decade may be
increased. To how many lines? The rule is that the
calculation should be started with the minimum num-

315

ber (one) per window. The number may then be incremented until there is no further improvement.
When the number of lines per decade is increased
the windows will necessarily overlap. In the overlap
regions the algorithm will then engage in unnecessary
calculations by going over the overlap regions twice.
These unnecessary calculations can be avoided. Let us
examine the situation in detail. Figure 3 plots the
negative inverse of the first logarithmic derivative of
the transient kernel, exp ( - t / r ) , defined by
Dtran ( t / c ) -

d
- exp ( - t / r )
d log t / r

= 2.303 ( t / z ) exp ( - t / z ) ,

(9)

as a function of log ( t / z ) for each of three neighboring spectrum lines. Window 1 is shown in the same
figure for comparison.
The span between the points of intersection of
Dtran(t/rk) with Dtran(t/rk-1) on the left, and
Dtran ( t / r k + l ) on the right, defines the Modeling Window. We denote the endpoints of the Modeling Window (Window 2) by tt and tu (for lower and upper,
respectively). Beyond these endpoints we enter the domains of the neighboring Windows 2. For this reason,
datum points for calculating the kth spectrum line
should be drawn from Window 2 when computing
more than one line per decade. The limits tt and t u
satisfy the equations
( t t / r k - 1 ) exp ( - t ) l / r k _ 1 = (tl/rK ) exp ( - tt/rk)

(lo)
and
( 6 / r k ) exp ( -- tu/r~) = (tu/rk + ~) exp ( -- tu/rk + i )

(11)
A little rearrangement yields

i.O
.8

t l - --rk'Ck--------~ ln rk
rk-- Tk_ 1 rk-1

and
tu-- r k + l r k lnrk+l
rk+l-rk
rk

.2
0

log (t/~)
Fig. 3. Definition of Window 2 (the Modeling Window)

(12)

(13)

If n is the number of spectrum lines per decade of


logt, then we have log(rx+l/rx) = 1/n and,
therefore,

tl-

2.303
(101/,
"ok
n
-1)

(4)

316

Rheologica Acta, Vol. 32, No. 3 (1993)

and

the imposition of a strain or a stress as a step function


of time such as, for example, the wave modulus or the
bulk compliance (Tschoegl, 1989; p. 515 ff).

2.303 x 101/n
t u --

n (101/n -- 1 )

r~.

(15)

Relaxation modulus
Values for n = 1 to n = 8 are tabulated below.
Table 1. Limits of Window 2
n

log tt/r ~

log tu/r k

log tl/zg

log t~/r k

1
2
3
4

-0.59
- 0.27
-0.18
-0.13

0.41
0.23
0.15
0.12

5
6
7
8

-0.10
- 0.08
-0.07
-0.06

0.10
0.08
0.07
0.06

i=N

G(t) = [Ge}+ ~ Giexp ( - t / r i )

(17)

i=1

The window becomes more symmetrical, and its


width decreases as the number of spectrum lines included within it increases. Accordingly,
lira Window 2 = 0

According to the theory of linear viscoelastic


behavior the shear relaxation modulus may be represented (Tschoegl, 1989; p. 121) by a Wiechert (or
generalized Maxwell) model as

(16)

n ---r ~

marks the transition from the discrete to the continuous form of the representation of G(t). We emphasize that W i n d o w 2 must contain at least one
datum point. In practice, this puts some limitation on
the width of Window2, i.e., on the number of
preselected spectrum lines one may choose.
Window 1 spans the region in which the exponential kernel function is at all effective in modeling the
relaxation modulus. Window 2, on the other hand,
demarcates the region over which this modeling is
most effective. Table 1 shows that the two windows
coincide for the choice of one spectrum line per
decade. We have chosen the limits of Window i as
- 0 . 6 and 0.4 instead of - 0 . 5 9 and 0.41 purely for
convenience in later calculations. The small difference
is without practical consequence.

The algorithm
The preceding theoretical discussion served to
clarify which datum points should be considered in
the evaluation of a particular spectrum line. We now
develop the detailed algorithm for the calculation of
the distribution of spectrum lines from experimental
responses to excitations applied as step functions of
time. We take the (shear) relaxation modulus first,
and follow this with the (shear) creep compliance. The
method is easily adapted to dealing with the tensile
relaxation modulus and with the tensile creep compliance or, for the matter, with any other response to

where "ci=rli/G i, and Gi and ~/i represent the


modulus and the viscosity of the ith Maxwell unit. G e
is the equilibrium modulus. The braces signify that
{Ge} = G e when the modulus describes an arrheodictic
material, and that (Ge} = 0 when the material is
rheodictic. 1) For convenience in the computations,
we normalize by the difference Gg-[Ge}, where Gg is
the instantaneous (or glassy) modulus. We remark
that the sum of the normalized spectra is always unity. Normalization leads to
i=N

g(t)={ge}+ ~ g i e x p ( - t / z i )

(18)

i=1

where g(t) is the normalized relaxation modulus, ge


is the normalized equilibrium modulus, and the gi's
represent the strengths of the delta functions in the
normalized discrete relaxation spectrum (Tschoegl,
1989; p. 227)
i=N
h(~-)=

~ gi'cit~('c-'ci).
i=1

(19)

We assume the source data to be available as a


discrete set of M datum points {G(tj)} where j =
1,2 . . . . . M. Each of these datum points can be normalized by the difference between the largest point,
G1, and the smallest point, GM, to yield the normalized set {~(tj)}. We can then express the modulus
alternatively by a discrete set of (normalized) spectrum lines, [~i}. In terms of these each datum point
becomes
i=N

~(tj) = ~(tM)+ ~ ~iexp ( - t j / r i )

(20)

i=1

1) The term rheodictic refers to a material showing


steady-state flow. Arrheodictic then denotes a material
which does not (Tschoegl, 1989; p. 93).

Emri and Tschoegl, Generating line spectra from experimental responses


We intend to determine, from the set of source
data, {~(tj)}, a set of spectrum lines, {~fl, which will
faithfully reproduce the modulus, G (t). As in Eq. (3),
we split the sum in Eq. (20) into three parts and write

317

where
i=k-1

~(tj) = g ( t j ) - - g ( t u ) -

gi

exp ( - tj/%)

i=m
i=N

i=k-1

g(tj) = ~(tM) +

--

giexp ( - tj/zi)

2
i=k+l

gi

exp (

tj/~ci)

(25)

i=l

+gk exp ( - ty/rk)


i=N

giexp(-tj/zi)+Aj

(21)

i=k+l

We have added the term Aj = A ~ + A ~ to allow us to


account for the experimental error in the source data,
A~, and for the approximation error, A i, which
arises because the calculated line spectrum is an approximation to the unknown true spectrum.
Two decades to the right of the location of any zi,
e x p ( - t / r i ) has already dropped to 3.7210 .44
which, to all intens and purposes, is zero (see Fig. 2).
We can therefore safely neglect the contributions of
all lines located two or more decades downscale from
r k. By beginning the first summation in Eq. (21) with
l<_m = k - 2 n - 1
instead of 1, we may thus reduce
the time which the algorithm needs to execute each
sweep. We therefore rewrite Eq. (21) as
i=k-1

~(tj)=g(tM)+

~,

Creep compliance

~iexp(-tj/zi)

We now turn to the creep compliance. For a Kelvin


(or generalized Voigt) model this becomes (Tschoegl,
1989; p. 123)

i=m

+ ge exp ( - tj/rk)
i=N

is the expression from which we obtain the strength of


the kth spectrum line.
We start the computation with the Nth, i.e.,
the farthest right, spectrum line, gN, because the
first sum in Eq. (25) then vanishes (cf. Eq. (4)). In the
first pass over the source data we set the ~j,
(i = m . . . . . k - 1), to zero. In succeeding sweeps, we
then substitute any newly found non-negative value
for the corresponding previous one and set any
negative value again to zero. The iteration is broken
off when the difference between the previously found
and the newly computed spectrum lines is smaller
than a preset error. We generally set the criterion for
terminating any of the required iterations as 0.0010/o
of the difference between the new and the previously
calculated value. Sharpening the criterion to 0.0001 7o
significantly increases the time required for the
calculations, but does not result in any significant
change in the values of the spectrum lines. This shows
that the algorithm converges.

i=N

giexp(-tj/ri)+Aj.

(22)

i=k+l

J(t)-{Ozt} = Jg+ ~ Ji[1 - e x p ( - t / r i ) ]


i=1

The sum of squares of Aj within Window 2 is given


by

i=N

= j~o I _ ~ Ji exp ( - t/zi) ,

(26)

i=1
J =~k, u

Ek =

Aj ,

(23)

j = Sk, l

where sk, l and Sk, u are the first and the last discrete
points in the window which belong to the kth spectrum line. Minimizing the error according to
OEk/O~ k = 0, leads to
J = Sk, u

~b(tj) exp ( - tj/zk)

g k - - :j = s/c' !
J =s~u

(24)
exp ( - 2 tj/rk)

J = slc, l

where Je is the equilibrium compliance, Je0 is the


steady-state compliance, Jg is the glassy compliance,
and q~f is the steady-flow fluidity, reciprocal of the
steady-flow viscosity, ~/f. The symbols in braces are
absent when the material is arrheodictic, and present
if it is rheodictic. Comparison with Eq. (17) shows
that the kernel function in the second of the two equations is identical with that of the relaxation modulus.
We therefore simply apply the algorithm to the creep
compliance in this form. Normalizing by the difference, J M - J 1 , we obtain equations which are formally identical with Eqs. (18) through (24). The
algorithm to be used for the creep compliance is thus

318

Rheologica Acta, Vol. 32, No. 3 (1993)

mathematically identical with that used for the relaxation modulus.

Results

We demonstrate the power of the algorithm first by


obtaining the singlet and doublet spectra which
characterize the standard arrheodictic model (threeparameter Maxwell modeO, and the standard rheodictic model (four-parameter model), respectively. We
follow this with two distributions of spectrum lines
associated with relaxation moduli modeled by
mathematical equations. We choose the latter f r o m
a m o n g those for which we also know the mathematical equation for the corresponding spectrum for
comparison.
The relaxation modulus of the standard arrheodictic model is given by
(27)

Garrh(t ) = G e + G exp ( - t / r ) .

We generated ten datum points per decade f r o m


Eq. (27) using Ge = 106, G = 109 N / m 2, and ~ = 10 -4
seconds in the interval -10_<logt_>0. Figure4
displays the modulus obtained by the algorithm,
Garrh(t).
The normalized results, gi, for the first, and for the
final, sweep are tabulated in Table 2. We have omitted the entries below l o g t = - 6 . 0 0 ,
and above
l o g t = - 2 . 0 0 because they were all zero. The gi obtained in the first sweep are shown in the second column. There are some smaller lines to the left and right
of log r = - 4 . 0 0 , where the spectrum line should lie.
In the final sweep, we obtained the expected singlet
spectrum shown in the last column. After multiplica-

8
7
Z

'

G,rra(l)

__w~5

,oltl

3
I

II

log I - seconds

Fig. 4. The relaxation moduli,


log Grheo(t), as functions of log t

log Garrh(t),

and

tion by the normalizing factor, G o - G e = G, we


recover the correct value of 109 N / m ~. The shape of
the spectra obtained in the initial sweeps depends, of
course, on the assumptions made for the location and
width of the two windows.
For the standard rheodictic model the relaxation
modulus becomes
Grheo (t) = Go exp ( - t/ro) + G1 exp ( - t/'c 1) .

(28)

It, too, is shown in Fig. 4. We again generated ten


datum points per decade, this time f r o m Eq. (28)
in the interval - 6 _ < l o g t _ < 3 , using G o = 106 ,
G1 = 109N/m 2, "Co= 1, and "q = 10-4seconds. The
results are shown in Table 3 where, as before, we left
out the lines at both ends of the calculated spectrum
since these again all vanish. After multiplication with
the normalizing factor, Gg = Go + GI, we recover the
original doublet spectrum.
In both these cases the location(s) of the relaxation
time(s) of the preselected set coincided with the location(s) of the "true" spectrum lines. When no
Table 2. Spectrum lines for the standard arrheodictic model
log t

1st Sweep

Final Sweep

- 6.00
- 5.50
-5.00
-4.50
-4.00
-3.50
- 3.00
-2.50
-2.00

0
0
0
0.0746
0.8930
0.0527
0.0007
0
0

0
0
0
0
1.0000
0
0
0
0

Table 3. Spectrum lines for the standard rheodictic model


log t

1st Sweep

Final Sweep

- 6.00
-5.50

0
0

0
0

- 5.00

0.0753
0.9011
0.0533
0.0007

0
0.9990
0
0

- 2.50

- 2.00

- 1.50
- 1.00

0
0

0
0

- 0.50
0.00
0.50

0.0001
0.0009
0.0001

0
0.999
0

1.00 "
1.50

0
0

0
0

2.00

4.50
4.00
3.50
3.00

10 -3

Emri and Tschoegl, Generating line spectra from experimental responses


preselected line coincides with an original spectrum
line, the algorithm must generate a composite line
spectrum to simulate the effect of the original line as
best it can. The composite spectrum will consist of
two or more lines. This results in a corresponding
broadening of the relaxation modulus reconstructed
according to Eq. (28) from the calculated distribution. It may be possible to improve the distribution by
increasing the number of preselected relaxation times
per logarithmic decade. The closer one of these comes
to the original relaxation time, the better will be the
calculated distribution. When coincidence occurs,
then we recover the relaxation modulus exactly.
To show this, we calculated the spectrum lines obtained from the standard arrheodictic model (see
Eq. (23)) with log ~ = - 4 . 2 , and the same moduli as
before. We preselected the relaxation times to lie at
log r . . . . - 5 , - 4 , - 3 . . . . . In each succeeding run
we increased the number of lines within each
logarithmic decade from 1 to a total of 5.
The results are tabulated in Table 4. With 1 to 4
line(s)/decade the algorithm generates doublet spectra
instead of the correct singlet spectrum. The two nonzero lines progressively approach one another. The
lines below l o g r = - 4 . 0 0 increase at the expense
of the line at log r = - 4 . 0 0 until, with 5 lines/decade, the expected singlet spectrum emerges at
log r = - 4.20. If the "original" line had been located
at log r = - 4 . 2 3 , say, we would still not have obtained the "true" spectrum.
The case of an "original" singlet spectrum is, of
course, highly unrealistic. In the general case, the observed relaxation modulus must be described with a
distribution of reasonably closely spaced spectrum
lines which will have no connection with any
"original", or "true" lines. We cannot even know if
such lines indeed exist. The question then arises: how
many lines per decade should the algorithm employ to

319

make possible the recovery of the original relaxation


modulus with an acceptable error? This error is simply the relative error derived from a comparison of the
original modulus data with those reconstructed from
the calculated distribution. The practical expedient is
to run the algorithm with an increasing number of
lines per decades until the desired relative error is attained. It must be borne in mind, however, that an increase in the number of spectrum lines per decade requires an increase in the number of datum points. If
no more datum points are available, W i n d o w 2
should be widened without, however, exceeding the
width of Window 1.
To allow us to gain insight into the working of the
algorithm and to show its resolving power, we have
hitherto discussed models containing one and two
Maxwell units. We now examine a Wiechert model
with a much larger number of units. The model we
have chosen simulates the monotone non-increasing
relaxation modulus of an entangled polymer. To obtain its parameters we read a set of Gi's and ri's from
the continuous relaxation spectrum generated from a
bimodal form of the Kobeko equation [(Tschoegl,
1989) p. 320], i.e., from [(Tschoegl, 1989) p. 493]
HB(r) = (Gg - Gp)('Co/7~72)1/2 exp ( - r 0 / r )
+ G p ( r l / r ) exp ( - r l / r )

(29)

where Gp is the pseudo-equilibrium (or plateau, or


entanglement) modulus. We used the values
Gg = 109, Gp = 106 N / m 2, r0 = 10 -4, and r~ = 106
seconds for the parameters of the equation. H s ( r ) is
plotted in Fig. 5. We read off the set [Gi, ril from this
curve in the interval from - l o g 5.49 to log 10.01 in
steps of 0.5 of a logarithmic unit. They were

9
Table 4. Effect of increasing the number of spectrum lines
per decade. Standard arrheodictic model
log r

1/dec

2/dec

- 5.00
4.80
- 4.75
- 4.67
- 4.60
-4.50
4.40
-4.33
-4.25
4.20
-4.00

0.3350
.
.
.
.
.
.
.
.
0.6650

.
-

3/dec

4/dec

5/dec

0
-

H~(~)

.
.
0.5259
.
.
0.6831
.
.
0 . 4 7 4 1 0.3169

z=o

0
0.8393

_2,,
'-8 '-6

1.0000
0.1607

-4 -2 0 2 4
log t, Iog z:- seconds

10

Fig. 5. Discrete distribution of spectrum lines and the relaxation modulus derived from it

320

Rheologica Acta, Vol. 32, No. 3 (1993)

deliberately displaced by 0.01 logarithmic units from


the location of the spectrum lines preselected for the
algorithm. Equation (29) thus simply served as an "inspiration". The procedure yielded the discrete spectrum

Hw(Z ) = ~ Gi~:,~(z- ri)


i

(30)

for the Wiechert model. The spectrum is shown in


Fig. 5 together with the relaxation modulus

Gw(t) = ~ Gi exp ( - t/zi)

(31)

i
derived from it.
We engaged the algorithm a second time to determine a second set of parameters {Gj, z]}, also in steps
of 0.5 logarithmic units, but in the interval - log 5 . 5 0
to log 10.0. Despite the deliberate displacement introduced betwen the two sets, the moduli generated
from them are virtually identical. They overlap to
such an extent that they cannot be resolved on the
scale of the plot. In this exercise, we have generated
two slightly different discrete distributions of relaxation times which, however, represented the same
underlying unknown distribution. Although the data
sets differed from one another, they generated moduii
which were virtually indistinguishable from each
other. This strongly supports the contention that the
algorithm is capable of producing a discrete set of
spectrum lines from which the experimental response
can be recovered faithfully.
It is another remarkable feature of the algorithm
that it will calculate nearly correct spectrum lines even
from only a portion of the input data. To show this,
we obtained the spectrum lines from the segment of
Gw(t) lying between log t = 3 and 7. This portion
constitutes an especially severe test. The filled rectangles in Fig. 5 mark the heights of the spectrum
lines obtained from this portion. The deviations at
both ends are due to truncation errors, i.e., the
absence of information outside of the selected portion. The slight differences in the magnitudes of the
lines of the full and of the partial sets can be traced
to the change in slope of the envelope. In the region
between log t = - 3 and 1, for example, where the
envelope is a straight line, the agreement between the
full and the partial set is virtually complete. The filled
circles represent the modulus calculated from this
spectrum. Despite the deviations between the two
spectra, the filled circles still overlap Gw(t).
Truncation errors do, of course, always occur
because we never have experimental information ex-

tending from t = 0 to t = ~ . The magnitude of the


deviations caused by this unavoidable truncation of
the available information depends primarily on how
close GI is to Gg, or GM is to G e (which, of course,
may be 0).

Discussion
In a forthcoming publication (Tschoegl and Emri,
to be submitted) we intend to compare the results obtained with our algorithm with those produced by
other methods. It might be useful, however, to contrast our algorithm briefly with the Multidata method
of Cost and Becker (1970), because it clearly
demonstrates the difficulties that may arise when
negative spectrum lines are generated. When applied
to the standard linear models the Multidata method
yields spectrum line distributions as good as those
generated by our algorithm provided that a line of the
predetermined set coincides with an "original" spectrum line. Figure 6 shows a plot of Garrh(t ) generated
from Eq. (6) with the same values for the moduli but
with log r 0 = - 4.01.
Because of the slight displacement, the spectrum
lines found by the algorithm were log G 1 = 0.0311 at
log t = - 4.50, and log G2 = 0.9689 at log t = - 4.00.
Nevertheless, the relaxation modulus calculated from
these two lines using Eq. (31) was identical with that
shown in Fig. 4 within the resolving power of the plot.
The Multidata method recovered a much more complex set of spectrum lines in which the modulus of
almost every second line of the spectrum was
negative. This resulted in the oscillations exhibited by
the curve marked A in Fig. 6. Removal of the negative
I

N8
!

"-7
6
5
-

/
i
-5

10g ~,0=-4"01
A
/
i
i
-4
-3
-2
10g t - seconds

t
a
-1

Fig. 6. Recalculation of log Garrh(t) as a function of log t;

a) from the singlet spectrum obtained by the algorithm, and


b) from the spectral distribution obtained by the Multidata
method

Emri and Tschoegl, Generating line spectra from experimental responses


lines f r o m the spectrum produced curve B. The
Multidata method will produce oscillations in the
modulus when, at a given point along the time scale,
the contribution of negative spectrum lines outweighs
the contribution of the positive lines. Since the contribution by the largest lines in the transition region
has effectively died out two decades upscale, these
oscillations are most likely to occur in the equilibrium
region of an arrheodictic material.
When we applied the Multidata method to Gw(t),
as given by Eq. (31), the spectrum was virtually identical with ours, except in the glassy region were both
positive and negative lines showed up where there
were no lines in the input spectrum, H w ( r ) , and
where our algorithm produced none. This, it should
be noted, has no noticeable effect on the reconstructed modulus because the contribution by positive
lines in this region compensates the contribution
which the negative lines make. In addition, the sum,
i Gi, is large enough in the glassy region to insure
that the positive contributions outweigh any remaining negative ones. On the right end of the spectrum we
now encountered no negative spectrum lines because
we were describing a rheodictic material with a monotone non-increasing modulus. The spectrum produced
by the Multidata method could conceivably be improved in a given case by a trial-and-error shifting of
the time scale. This, however, is evidently unsatisfactory as a routine procedure.
We conclude that in applying the Multidata method
one m a y routinely encounter unwanted lines, some of
which will necessarily by negative. Even when this has
no noticeable effect on the reconstructed modulus or
compliance, it must be expected to vitiate any effort
to convert Gi's into Ji's and vice versa.

Acknowledgements
The authors gratefully acknowledge support of this work
by the Slovene Ministry of Science under Grant
P2-1131-782/91, and partial support by the California
Institute of Technology.

321

method and its application to creep data. Macromolecules


24:310- 314
Elster C, Honercamp J (1992) The role of the error model
in the determination of the relaxation time spectrum. J
Rheol 36:911 - 927
Elster C, Honercamp J, Weese J (1991) Using regularization
methods for the determination of relaxation and retardation spectra of polymeric liquids. Rheol Acta 32:262 - 174
Emri I, Tschoegl NW (1993 b) Generating line spectra from
experimental responses. IV. Application to experimental
data. To be submitted to Rheol Acta
Honercamp J (1989) Ill-posed problems in rheology. Rheol
Acta 28:363 - 371
Honercamp J, Weese J (1989) Determination of the relaxation spectrum by a regularization method. Macromolecules 22:4372-4377
Janzen J, Scanlan J (1994) Practical implementation of differential operator methods for determining continuous
distributions of relaxation times from linear viscoelastic
data. Submitted to Rheol Acta
Kamath VM, Mackley MR (1989) The determination of
polymer relaxation moduli and memory functions using
integral transforms. J Non-Newt Fluid Mech 32:1199
Laun HM (1986) Prediction of elastic strains of polymer
melts in shear and elongation. J Rheol 30:459
Orbey N, Dealy JM (1991) Determination of the relaxation
spectrum from oscillatory shear data. J Rheol 35:
1035 - 1049
Schapery RA (1962) Approximation methods of transform
inversion for viscoelastic stress analysis. Proc Fourth US
Nat Congr Appl Mech 2:1075
Tobolsky AV, Murakami K (1959) Existence of a sharply
defined maximum relaxation time for monodisperse
polystyrene. J Polymer Sci 40:443
Tschoegl NW (1989) The phenomenological theory of linear
viscoelastic behavior. Springer, Berlin
Tschoegl NW, Emri I (1992) Generating line spectra from
experimental responses. III. Interconversion between
relaxation and retardation behaviour. Intern J Polym
Mater 18:117- 127
Tschoegl NW, Emri I (1993 a) Generating line spectra from
experimental responses. II. Storage and loss functions.
Rheol Acta 32:322-327
Tschoegl NW, Emri I (1994) Comparison of various
algorithms for generating spectral functions from experimental responses. To be submitted to Rheol Acta
(Received August 24, 1992;
accepted April 1, 1993)
Correspondence to:

References
Baumgaertel M, Winter HH (1989) Determination of
discrete relaxation and retardation spectra from dynamic
mechanical data. Rheol Acta 28:510-520
Cost TL, Becker EB (1970) A multidata method of approximate Laplace transform inversion. Intern J Num
Methods in Engg 2:207
Elster C, Honercamp J (1991) Modified maximum entropy

Prof. Igor Emri


Mechanical Engineering Department
University of Ljubljana
Murnikova 2
61000 Ljubljana, Slovenia
Prof. N.W. Tschoegl
California Institute of Technology
1201 East California Boulevard
Pasadena, CA 91125, USA

S-ar putea să vă placă și