You are on page 1of 10

IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 25, NO.

4, OCTOBER 2010

2997

Strain Measurements on ACSR Conductors During


Fatigue Tests IIStress Fatigue Indicators
Sylvain Goudreau, Frdric Lvesque, Alain Cardou, and Louis Cloutier

AbstractFatigue strength of overhead conductors of different


types is often presented on the same diagram, a so-called S-N diagram, the vertical axis being generally taken as the alternating
stress amplitude. In this paper, it is shown that these stress levels
are merely fatigue indicators with inherent limits, and should be
used accordingly. Their practical interest is to allow a regrouping
of the fatigue results of different types of conductors. It will be
shown that these indicators have some inherent limits in their definition and can only be used according to their intended purpose.
Index TermsAluminium conductor, fatigue, strain measurement.

Fig. 1. Standing-wave vibration.

I. INTRODUCTION

N THE FIRST paper [1], measured strains at several cross


sections on the outer layer uppermost and lowermost wires
for two conductors under cyclic bending were reported. The experimental method followed and the mathematical treatment applied onto the experimental data was also summarized. In the
last section of this paper, it was shown that the two alternating
and
, which are used to express the
stresses PS
fatigue severity undergone by a vibrating conductor fitted on
a short metallic suspension clamp, are rather different from the
alternating stresses that might be computed from the strain measured on the vibrating conductor.
In this paper, first, the development leading to equations
and
will be summarized and,
yielding PS
second, the influence of the parameters governing these expressions will be examined. It must indeed be noted that neither of
them takes the geometric characteristics of the supports into account. Finally, a review of several technical reports and papers
reporting measured strains will be presented and a discussion
on how these measured strains correlate with the calculated PS
and
values will follow.

Fig. 2. Deflected shape in the clamp region.

steps, this development is identical to the one found in [2] and


[3]. These two stresses are used as fatigue severity indicators in
order to regroup all fatigue results of different conductors in a
single S-N diagram. To begin, it is assumed that some distance
away from a fixed support, a vibrating conductor takes the shape
of a sine wave (Figs. 1 and 2)
(1)

II. STRESS FATIGUE INDICATORS


and
The development of the two alternating stresses PS
is summarized in what follows; except for a few

Manuscript received October 02, 2009. Current version published September


22, 2010. Paper no. TPWRD-00739-2009.
S. Goudreau, A. Cardou, and L. Cloutier are with the Department of Mechanical Engineering and the GREMCA Laboratory (Groupe de REcherche
en Mcanique des Conducteurs Ariens). Universit Laval, Qubec City, QC
G1V 0A6, Canada (e-mail: Sylvain.Goudreau@gmc.ulaval.ca; acardou@gmc.
ulaval.ca; louisjcloutier@sympatico.ca).
F. Lvesque is with the Department of Civil Engineering, Universit de Sherbrooke, Sherbrooke, QC J1K 2R1, Canada (e-mail: frederic. levesque@usherbrooke.ca).
Digital Object Identifier 10.1109/TPWRD.2010.2042083

where
is the antinode amplitude of the sine-shaped deflecis the distance from the
tion curve, the wavelength, and
fixed support beyond which the sine-shaped deflection curve
is defined. The fixed support is assumed to be a square-faced
bushing.
The bending curvature of the conductor is assumed to be
given by the usual solid beam equation (Fig. 3)
(2)
where is the departure of conductor centerline (Fig. 3) from
in the vicinity of the bushing. It will
the sine-shaped loop

0885-8977/$26.00 2010 IEEE

2998

IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 25, NO. 4, OCTOBER 2010

where is the wire diameter on the outer layer of the conductor


(in fact, it can be the diameter of any wire of the conductor),
Youngs modulus of aluminium,
the conductor minimum
bending rigidity
, which is the sum of all individual wires
bending rigidity.
Depending on the engineering needs, may be expressed in
terms of the vibration parameters: frequency and the antinode
amplitude
. Using the velocity of traveling waves on a taut
string,
(10)
Fig. 3. Deviation of the conductor centerline from the sine-shaped loop near
the fixed end.

may be expressed as
(11)

be assumed that the angular amplitude


in Fig. 2 is small so
that the coordinate is assumed to be equal to . is the local
is the equivalent local bending rigidity
bending moment and
and assuming that the usual
of the conductor. Since
beam bending equation holds, the preceding equation yields

Another expression for


can be derived from the local field
(Fig. 2) near the bushing mouth. Assuming small deflections,
the position of the conductor centerline relative to the axis
is
(12)

(3)
From (6)
With the boundary condition that approaches zero for large
, the solution to the differential equation is given by
(4)
The slope of the conductor relative to the

(13)
and, combined with (4), (12) becomes
(14)

axis is
(5)

Finally, factoring out

in (14) and using (6) yields


(15)

and at
0,
is equal to
(Fig. 2). The curvature
of the conductor at the mouth of the square-faced bushing is
(6)
is assumed to be equal
As a first approximation, the angle
of the sine-shaped deflection
to the maximum value of
curve and it is given by
(7)

where the standard position for evaluating


is
89
is obtained
mm from the bushing mouth. The value of
by measuring the so-called peak-peak bending amplitude
. For a short metallic suspension clamp, the
standard position for measuring is 89 mm from the last point
of contact (LPC) between the conductor and the bed of the
clamp.
Combining (9a) and (15) yields another expression for

The curvature of the conductor at the mouth of the square-faced


bushing becomes
(8)
It is assumed that each wire of the conductor bends about its
own neutral axis and the alternating stress amplitude
at the
mouth of the square-faced bushing is expressed as
(9a)
(9b)

(16)
This last equation is known as the Poffenberger & Swart stress
(PS ).
III. PARAMETERS INFLUENCING THE COMPUTED
Both (11) and (16) seem to indicate that stresses are directly
proportional to wire diameter . So it would be interesting to
check how sensitive these equations are with respect to .
Assume: 1) that a multilayer conductor is made of wires
of the same material (aluminium), which is the case in ACAR,
AAC, AAAC conductors; 2) that all wires are of the same diameter ; 3) that all wires act independently in bending (thus

GOUDREAU et al.: STRAIN MEASUREMENTS ON ACSR CONDUCTORS DURING FATIGUE TESTS II

2999

TABLE I
GEOMETRICAL AND MECHANICAL PROPERTIES OF ACSR CONDUCTORS

yielding the minimum rigidity


becomes

). In the case of (11),

(17)
where
is the density of aluminium kg/m .
It is found that no transverse dimension of the wire section
appears in the stress equation meaning that the last expression is
independent of wire diameter and, thus, the equation of stress
is proportional to a constant which depends only on material
properties. In the case of aluminium
(18)
In fact, for aluminium-conductor steel-reinforced (ACSR)
multilayer conductors, [2] and [3] indicate that the constant
may vary from 0.171 to 0.200 and in the same
reports, it is suggested to use a mean value of 0.186 (MPa
is
s/mm). It should be remembered that the computation of
based on the assumption of independent bending of wires.
Now, with the other limit hypothesis that all wires act as
if they were welded together (no slip), the bending rigidity is
and is calculated by considering the conductor as a solid
beam. In this case, the overall diameter of the conductor is
used instead of the wire diameter in (9a). The resulting computation of
of conductors tested to build the S-N
diagram shown in Figs. 4 and 5 is summarized in Table I. The
differs by a factor of, at most, 15%
relationship versus
. It should be noted, however,
from the one computed with
that the ACSR conductors in this table do not exactly meet the
previous hypotheses, as they include a 7 steel-wire core.
is used as a fatigue severity
The alternating stress
indicator in order to regroup all fatigue results of different conductors in a single S-N diagram. In Figs. 4 and 5, most data
points are from [2, Figs. 3.213a and b], with some additional
data points [4], [5]. Each data point is computed with its own
. In Figs. 4 and 5,
proportionality factor (Table I) by using

Fig. 4. Fatigue tests of a two-layer ACSR. This is from [2, Fig. 3.213a] updated with some additional data points.

Fig. 5. Fatigue tests of a three-layer ACSR. This is from [2, Fig. 3.213b],
updated with some additional data points.

some data points deviate slightly from those shown in [2, Figs.
3.213a and 3.213b] since those data coming from [3] computed with a average value of the proportionality factor of 0.186

3000

IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 25, NO. 4, OCTOBER 2010

were used again in [2] and some raw data from the GREMCA
laboratory have been slightly adjusted. Also, some data points of
the ACSR Drake conductor [10] have been modified according
to [4]. Looking at Figs. 4 and 5, the fatigue results expressed
as , which are practically directly proportional to a value of
, are very consistent. As reported in [2], the S-N
diagram for three-layer ACSR conductors differs from the one
obtained for the two-layer ACSR conductors; their endurance
limits , defined as the highest alternating stress amplitude
with a zero wire break at 500 Mc, are about 22 MPa and somewhat less than 30 MPa, respectively. These S-N diagrams are
for ACSR conductors fitted in short commercial metallic suspensions clamps and in metallic bell mouth clamps.
The fact that (11) is practically independent of diameter and
of the bending rigidity hypothesis seems to indicate that drawing
taken as the ordinate is as good as
the S-N diagram with
(11).
using
By comparison, it will now be shown that the value of
yielded by (16) is not directly proportional to the wire diameter.
The mathematical development is based on the same assumpbe the normal traction stress computed
tions as before. Let
according to (19) (lay angles are neglected)

Fig. 6. Fatigue tests of two-layer ACSR. From Fig. 3.224 [2] with some additional data points.

(19)
and

becomes
(20)

Assuming that in the denominator of (16) with the term


, the following expression is obtained:
(21)

It is possible to deduce approximately the influence of wire diameter using (21). It appears that for a given static traction stress
, the alternating bending stress
is not proportional to wire
diameter .
In Figs. 6 and 7, while most data points have already been reported in [2, Figs. 3.224 and 3.225], there are some additional
data points [4], [5]. Considering all of the tested ACSR conductors (with the associated testing conditions) in Figs. 6 and 7, the
computed by using (16) varies from
corresponding ratio
27.2 to 46.3 (MPa/mm) (Table I). Using (21),
differs by,
at most, 6% from the one computed using (16).
, it is now assumed that the prevailing
Instead of using
and that the distance from the neutral
bending rigidity is
axis to the outer fiber is
. In this case,
is at least 2.6
(Table I).
times the value computed with
Comparing Figs. 6 and 7, one sees that the S-N diagram for
three-layer ACSR conductors differs from the one obtained for
the two-layer ACSR conductors; the vertical axis is the altercomputed by using
. The endurance
nating stress
for three-layer ACSR conductors is about 8.5 MPa.
limit
Howewer, for the two-layer conductor case, it is noted that at
the 15-MPa level, the number of cycles of the first wire break

Fig. 7. Fatigue tests of three-layer ACSR. From Fig. 3.225 [2] with some
additional data points.

varies from 17 Mc to 150 Mc and because of a lack of fatigue


test results in the range of 10 to 15 MPa, the corresponding
endurance limit cannot be clearly determined yet. These S-N
diagrams are for ACSR conductors fitted in short commercial
metallic suspension clamps and in metallic bell mouth clamps.
It should be noted that according to the macroscopic parameter
) used to compute
leads to different values of
( or
.
Using (20), (9b) yields
(22)
To obtain the fatigue S-N curve of a conductor-clamp
system, [2] and [6] recommend that fatigue tests be carried
out by using a resonance test bench. With such a bench, the
active length of the conductor is a multiple of the loop length
of the vibrating conductor. For a given loop length and a given
(or a given peak-peak amplitude
antinode amplitude
depending of the controlled parameter),
bending deflection
and
vary differently
the alternating stresses
to the traction force. The alternating stress , computed by

GOUDREAU et al.: STRAIN MEASUREMENTS ON ACSR CONDUCTORS DURING FATIGUE TESTS II

3001

TABLE II
EXPERIMENTAL DATA

All data were collected from original papers. Regression lines through experimental data were computed. Data appearing in [2] and [3] are indicated by .
There are some minor differences between the computed ratio in [2] and the ratio given in [3].
Experimental conditions:
obtained from experimental data f , y
;
1) regression line through five data points; strain gauges on wires of the outer layer; f y
obtained from , H and through (7) and (10);
2) regression line through two data points; strain gauges on wires of the outer layer; f y
3) regression line through 16 data points; the 16 data points come from curves in the original paper; curves obtained from strain gauge data measured on the
deduced from , H and through (7) and (10);
two uppermost and the two lowermost wires of the outer layer; f y
obtained from experimental data y
, , H ,
4) regression line through 10 and more data points; strain gauges on wires of the outer layer; f y
and through (10);
obtained from experimental data f and y
;
5) mean of four data points; one strain gauge on the two adjacent uppermost wires of the outer layer; f y
obtained from experimental data f and y
;
6) mean of two data points; one strain gauge on the two adjacent uppermost wires of the outer layer; f y
7) regression line through 4 data points; the nearest strain gauge from the Keeper Edge (KE) on the uppermost wire of the outer layer (Drake, wire 15;
Bersfort, wire 17) and the nearest strain gauge from the Last Point of Contact (LPC) on the lowermost wire of the outer layer (Drake, wire 113; Bersfort,
obtained from experimental data f and y
. In the calculation of the theoretical  f y
, EI
is computed with 4 aluminium
wire 17); f y
wires removed from the outer layer and m is computed with no aluminium wire removed.

using (22), will vary according to the square root of the traction
force and the variation of the alternating stress , computed by
using (21), is smaller. For example, consider the AAAC Aster
570. It is made of 61 wires (wire dia. 3.45 mm) and its construction is 24/18/12/6/1. Doubling the traction force from 20%
(22) by a factor
RTS to 40% RTS multiplies the ratio
. The ratio
given by (16) and (21) changes by a factor
of 1.265 and of 1.245, respectively. Sepp [7] and Poffenberger
and Komenda (discussion in [8]) reported that the strain measured in the clamp vicinity of a conductor varies as the square
root of the applied tension force.
It seems difficult to reconcile the two approaches based on
(11) and (16) since the traction force variation and the bending
rigidity variation affect the alternating stress differently. Obviously, the difference between (9b) and (16) comes from the

way the curvature at


inating equation is (6)

0 is computed. In both cases, the orig-

(23)
In (9b),
is obtained from the sine-shaped deflection curve
of the vibrating conductor through the following equation:
(24)
while in (16),
comes from (13) and it is expressed in term
of the amplitude A of the deflection curve. The value A is deter-

3002

IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 25, NO. 4, OCTOBER 2010

mined by measuring the peak-peak displacement at the position

(25)
It is easy to experimentally obtain an empirical relationship
between
and
at a given frequency of excitation but
and it does not provide
this does not give the exact value of
which position the equal slope between the deflection curve of
the local field near the bushing and the sine-shaped deflection
curve of vibrating conductor occurs. On the other hand, it should
be remembered that (23) and (25) are based on the assumption
of a uniform bending rigidity along the length of the conductor.
In fact, the bending rigidity is probably high (in the order of
) at a portion of the deflection curve very close to the fixed
end due to the clamping action of the bolting of the two halves
of the square-faced bushing (or of the bolting of the keeper for a
conductor-suspension clamp assembly) and in the asymptotic
portion of the deflection curve. Elsewhere, between these two
regions, the bending rigidity may be much lower, as slip between
wires is easier.

(central alternating) function of bending amplitude ACSR Drake


Fig. 8. "
with suspension clamp.

and in order to compare conductors of different sizes, the following ratio is used:
(27)

IV. EXPERIMENTAL DATA


A. Alternating Stress

Based on

Table II presents a summary of the experimental data reported


by different laboratories in order to compare the measured alternating strain (transformed into corresponding alternating stress)
to the stress calculated by using (11). Table II shows that with
only three exceptions, the data from tests with the square-faced
bushing follow more closely (11) than do data from tests performed using the short commercial suspension clamp. As already pointed out, (11) does not take into account the geometry
of the clamp and looking at the results of McGill and Ramey
([9], [10]), the radius of curvature of the generic clamp seems
to significantly affect the measured alternating strains. Also, in
the case of a short suspension clamp, strain gauges cannot be localized as close to the edge as with a square-faced bushing since
the exit of the keeper and of the bed of suspension clamp have
a relieving curvature which can interfere with the strain gauges
when glued too closely to the support.
The experimental data show a large scatter. As reported in
[2], Claren & Diana [8] found that strain measurements varied
a lot from one wire to the next ( 30%). Sepp [11] reported a
deviation of the measured dynamic strain value from the computed average of the order of 25%. Looking at [9] and [10], one
finds a variation of about 20% with respect to the mean value.
B. Alternating Stress

Based on

Equation (16) will be modified in order to facilitate the comparison between experimental and theoretical results; to compare measured strains, (16) is modified as follows:
(26)

measured at the uppermost poFig. 8 shows the strains


sition of the ACSR Drake conductor fitted in the conventional
suspension clamp reported by Poffenberger and Swart [13] and
in the generic clamps used by McGill and Ramey [9], [10]. The
generic clamps have a radius of curvature of 6 in. and 12 in.
with (deep bed) or without (shallow bed) a bed well fitted
to the conductor, and the sag angle is 11 . The straight lines correspond to the linear theory in (26) computed by using
at three traction force levels. The strain values shown for the
-axis interval [0.9 mm, 1.3 mm] [10] in Fig. 8 were measured on two adjacent wires and as close as technically feasible
from the keeper edge (KE). It can be noticed, even with the large
scatter for a given generic clamp, that increasing the radius of
curvature decreased the alternating strain
and, except for
one case, that the theoretical straight line does not provide a
good fit for these experimental data.
Fig. 9 is an enlarged view of Fig. 8 in the low amplitude region. Poffenberger and Swart [13] explained the nonlinearity of
the experimental data by changes in flexural rigidity caused by
interstrand slippage at higher displacements. The difference in
behavior at 15% and 25% RTS was associated with a speculative possibility that the last point of contact between the conductor and the conventional suspension clamp may have shifted
slightly in reducing the tension. If the last point of contact was
closer to the support center line, the corresponding lengthening
of the deflection arm and shifting the gauge would both result
in lower strain values. The two extreme points (partially filled
squares) on each experimental data set are two experimental
points used by Poffenberger and Swart [13] to build Fig. 11.
Fig. 10 shows experimental data from an IEEE Committee
report [14]. It shows measured strain values of an ACSR 583.2
kcmils 18/7 conductor fitted in an articulated suspension clamp
under a traction force of 30% RTS. The sag angle (determined

GOUDREAU et al.: STRAIN MEASUREMENTS ON ACSR CONDUCTORS DURING FATIGUE TESTS II

Fig. 9. "
(central alternating) versus bending amplitude ACSR Drake with
the suspension clamp.

Fig. 10. " (central alternating) versus bending amplitude 583.2-kcmil ACSR
conductor with a suspension clamp.

by using the ratio of the vertical load to the tension load) is approximately 4.6 . The peak-peak bending amplitude is measured relative to the clamp. Strain gauges no. 2, 4, and 5 were
glued on the three uppermost wires on the conductor at the last
point of contact between the top of conductor and suspension
clamp. Gauge no. 3 was located on the same wire as no.4 at 12.7
mm from that gauge, on the clamp side. Gauge no. 1 was located
on the gauge no. 2 wire at 12.7 mm from that gauge, on the conductor span side. As seen in Fig. 10, the scatter of strain values
can be high; for example, at
0.5 mm, the strain amplitude
given by gauge no. 4 is 35% higher than the mean value obtained
from the remaining four gauges. The upper data point (partially
filled square) corresponds to an experimental point used by Poffenberger and Swart [13] to draw Fig. 11.
Fig. 11 shows the experimental data for seven conductors
reported by Poffenberger and Swart [13] and those obtained
for seven conductors by Claren and Diana [8]. Each point corresponds to the maximum strain value at the maximum displacement obtained in any set of data, where interstrand binding
would be at a minimum. The data (six of seven conductors) reported in [13] come from unpublished reports by Rulhman and
Swart and by Edwards and Boyd at the publication time of [13].
The data from Rulhman and Swart and Edwards and Boyd reports are used in [14] and it is reported that tests carried on
unarticulated clamps with various mouth radii showed that the
strain-bending-amplitude relationship was constant with radii

3003

Fig. 11. Strain/displacement factor versus stiffness parameter (exp. points


computed by using EImin).

ranging from 1.6 mm (1/16 in) to 152 mm (6 in). The curve


drawn in Fig. 11 corresponds to the theoretical curve computed
. The solid symbols correspond to data
with (27) using
where the peak-peak bending amplitude
exceeds 0.25 mm
(0.010 in.) and the clear symbols correspond to less than 0.25
mm. The Claren and Diana data [8] were obtained for seven conductors fitted with a square-faced aluminum bushing at a zero
sag angle. Unfortunately, for those data reported in [13], the
exact boundary conditions (sag angle, geometry of the clamps)
were not given for each experimental point. Poffenberger and
Komenda pointed out that the scatter of the Claren & Diana data
[8] may be attributed to
not exceeding 0.25 mm (discussion
in [8]) and that using data obtained with amplitudes
larger
than 0.25 mm would show a better fit with the theoretical curve
based on
. A. T. Edwards (discussion in [8]) referring to
work reported in [14] mentioned that it was impossible to predict where the maximum strain would occur. Thus, it was found
to be necessary to install strain gauges on the top three strands
at the clamp just where the conductor leaves the clamp, and just
inside and outside the mouth of the clamp. The maximum strain
was chosen from the five separate strain amplitude readings at
and about the clamp.
Komenda and Swart [15] did perform a statistical analysis of
39 different laboratory tests and ten data sets measured at different field locations. It resulted in 49 different data groups; each
of points (only
data group contains a large number
three data groups have
). Each experimental point is
transformed according to
. For each data group,
a mean value
is computed and, finally, using
from (26),
is computed as
. This procedure enables
mapping the nonlinear behavior of data into a straight-line theoretical curve. Taking into account the 49 data groups, they computed a weighted average
and a standard deviation
0.246 using the weight
of each data group.
The corresponding confidence band is equal to [0.888, 1.077]
(Table III). They concluded that, on average, (26) is a reasonable approximation of the relationship between strain
and
.
Unfortunately, no information regarding the type of support
corresponding to each data group was reported in [15]. It was
mentioned in the paper that the strain gauges were located at
the point of maximum strain for the support configuration being

3004

IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 25, NO. 4, OCTOBER 2010

TABLE III
ANALYSIS RESULTS OF THE DATA REPORTED BY KOMENDA AND SWART [15]

studied but no specific information was given on how it was accomplished. All data were grouped together without regarding
the geometrical characteristics of the support (such as a squarefaced bushing) or a short commercial suspension clamp. The
prediction band (Table III), computed from the analysis based
on all data, gives an interval which probably characterizes the
effect of the geometrical parameters of the support, of the position where the strain gauges are glued, and of the variation from
one wire to the adjacent one if single strain gauge monitoring
is used; these effects can also be appreciated by the large variation of individual values of shown in the column identified as
.
Now, the precise mathematical meaning of the terms confidence and prediction band should be recalled. If 100 samples
from a population were drawn and a 99% confidence band of the
mean was computed for each sample, then 99% of these confidence bands would contain the true mean of the population. In
practice, usually one sample is drawn and one confidence band
is computed, so it cannot be said that this single confidence band
necessarily contains the true mean of the population. But it can
be said that with 99% confidence, the true mean of the population is contained in this single computed confidence band. This
last statement implies a notion of frequency: we do not know
whether this single confidence band contains the true mean of
the population but we do know that the method used to compute
the confidence band yields a true statement in 99% of the cases.
On the other hand, the prediction band shown in Table III gives
the interval in which the value of a single future observation is
expected to be in 99% of the cases.
Looking at the other results of Table III, which were obtained
following the same computational approach, several points
ought to be emphasized.
1) The field data probably correspond to a short metallic conventional suspension clamp and
is lower than 0.983.
2) Looking at the laboratory data of the ACSR Pheasant, the
resulting
is much less than 1.0 and the 99% confidence band does not include the value of
1; this indicates probably that
significantly departs from 1.0. Is
this behavior related to a specific conductor-support system
or to specific characteristics of the
measuring system?
(Is the
value identical in each data group?)
3) Considering the laboratory data of ACSR Drake, which has
a large number of data groups
22), the resulting
is higher than 1.0, and the 99% confidence band does not

1. The same question as in the


include the value of
precedent point arises.
A more refined statistical analysis should take into account
the clamp parameters and the strain gauge location; this analysis
should be based on data obtained with measured at the stanvalue (89 mm). Unfortunately, in [15], the
value
dard
used for each data group is not reported.
Scanlan and Swart [16] did test an ACSR Pheasant fitted
at both ends of the vibrating span with what seems to be
square-faced bushings. A zero sag angle and a rather low
traction force of 14.3% RTS were imposed. Spanwise dynamic
variation along the vibrating conductor was determined
by measuring with optical means the peak-peak transverse
displacements at 15 positions from the farthest fixed end from
the shaker inward to midspan. Several tests were carried out
at different values of midspan deflection. Using the classical
beam bending equation, dynamic bending stiffness was then
numerically determined along the 30-in-long half span. Effective dynamic
values as high as 56%
were found at
the fixed end with an average value of 42%
. The general
trend of the dynamic effective
was found to decrease at
positions away from the fixed ends.
Fig. 12 is obtained by grouping the data of [1], [8], [10], and
those reported in [13]. The data shown are those where
is
greater than 0.25 mm (0.010 in.) and each point corresponds
to the maximum strain value at the maximum displacement obtained in any set of data. The curve drawn corresponds to the
theoretical PS
(27) computed by using
(or
according to [15]). It can be noticed that the experimental data of
[1] and [10], which are associated with clamps that differ greatly
from a square-faced bushing, depart significantly from the theoretical curve which shows that clamp characteristics do influence strain value
.
In order to show the scatter of strain measured on adjacent
wires, strain data measured by McGill and Ramey ([9], [10])
on an ACSR Drake conductor are compared in Table IV with
strain values predicted by (26). Strain gauges are glued on the
two uppermost wires of the outer layer 6.35 mm away from the
KE. The mean of all measured strains corresponds to 54% of
the theoretical value. A scatter as high as 20% compared to
the mean of the two measured adjacent wires was found in each
test. As previously mentioned, this scatter was also reported in
[8] and can be seen in Figs. 10 and 12.

GOUDREAU et al.: STRAIN MEASUREMENTS ON ACSR CONDUCTORS DURING FATIGUE TESTS II

3005

TABLE IV
EXPERIMENTAL DATA FROM [10]

Assumed value: 1.295 mm (0.051 in.). The Y value reported in [10] for this test is 0.38 mm (0.015 in.). Comparing the strain values of this test with the strain

values of the Y

= 1.09-mm (0.043 in.) test (line above in Table), it seems that the Y

value (0.015 in.) is very low. All tests were carried out under the

same testing conditions and should give approximately the same values of Y for the 12-in. radius clamp. Checking in [9] and [10], the reported values of

computed stresses, the ratio of these computed stresses is 1.1837 which is approximately equal to the following ratio of Y values 0.051 in./0.43 in. Thus, the

= 284.5 mm/s (f = 32 Hz; double amplitude mid-loop = 17.8


= 11 deg.), 28% RTS, x = 89 mm. The strain data of this test are used in Figs. 8 and 12 with the assumed value of Y = 1.295 mm.

value of 0.015 in. seems to be a typographical error. The testing conditions were fy
mm), sag angle

The strain values of this test are not used in the computed mean of 54%.

Fig. 12. Strain/displacement factor versus stiffness parameter (exp. points


computed using EImin).

V. CONCLUSION
Section II gives the theoretical development of an alternating
based either on vibration parameter or on the bending
stress
and
. In
amplitude, which have been noted
Section III, it is shown that it is difficult to reconcile both stress
values because the traction stress influences each value according differently and because the use of the two limit bending
and
) has a much stronger influence on
rigidities (
value than on the value
. In Sections IV-A
the PS
and B, the correlation between the reported experimental data
was examined. It was found that the
and the theoretical
scatter of experimental strain values is large and that in many
cases, the correlation between experimental strains and theoretis weak.
ical
Figs. 47 are the fatigue S-N diagrams for two-layer and
three-layer ACSR conductors fitted on short commercial
metallic suspensions clamps and on metallic bell mouth
clamps. In these diagrams, the vertical axis is the alternating
or
. The two-layer
stress amplitude

and the three-layer ACSR conductors show a different pattern


with respect to the so-called Endurance Limit (which is
often defined as the highest amplitude with a zero wire break
after 500 million cycles). As reported in [2], more fatigue tests
are needed to better characterize the difference between the
two- and three-layer ACSR conductor endurance limit. In [2]
and [3], selecting the endurance limit for all multilayer ACSR
conductors based on the three-layer conductor fatigue diagram
either at
equal to 22 MPa or
equal to 8.5
MPa is suggested.
values by the aluminium
Dividing one or the other of
, one obtains a characteristic strain level
Youngs modulus
which may be compared with the strain measured on a conductor fitted to a particular suspension clamp. Using this
comparison to determine its specific fatigue performance or
the severity of vibration at the clamp is misleading because, in
data do not show a strong
many cases, the experimental
correlation to both theoretical models for . In fact, because
the endurance limits of three-layer ACSR conductors have
different values depending on the way they are determined, it is
equations on which these two models
most probable that the
are based do not contain the fundamental parameters which
are necessary to characterize the fretting fatigue phenomenon
involved in wire breaks. It is, of course, recognized that the
original intent of defining the alternating stress was to have a
way of regrouping fatigue results for different conductor types
and test conditions based on the use of macroscopic parameter
or ), thus eliminating the inherent scatter arising
(
from strains measured on an individual wire.
REFERENCES
[1] F. Lvesque, S. Goudreau, A. Cardou, and L. Cloutier, Strain measurements on conductors acsr conductors during fatigue tests IExperimental method and data, IEEE Trans. Power Del., vol. 25, no. 4,
pp. 28252834.
[2] L. Cloutier, S. Goudreau, and A. Cardou, Chapter 3: Fatigue of overhead conductors, in EPRI Transmission Line Reference Book: WindInduced Conductor Motion. Palo Alto, CA: EPRI, 2006.

3006

[3] C. B. Rawlins, Chap. 2: Fatigue of overhead conductors, in Transmission Line Reference BookWind-Induced Conductor Motion. Palo
Alto, CA: EPRI, 1979, EPRI Res. Project 792.
[4] G. E. Ramey, Conductor fatigue life research Jul. 1981, EL-1946,
EPRI Res. Project 1278-1, Final Rep.
[5] GREMCA, Fatigue tests on the 48/7 Bersfort acsr conductor at Yb
varying from 0.3 mm to 0.35 mm (in French) Dept. Mechan. Eng.,
Laval Univ., Quebec City, QC, Canada, Tech. Rep. SM-2007-01, 2007.
[6] CIGRE Task Force B2.11.07, Fatigue endurance capability of conductor/clamp systemsUpdate of present knowledge, in Confrence
Internationale des Grands Rseaux lectriques, 2006, Tech. brochure
no TBD.
[7] T. O. Sepp, Self-damping measurements and energy balance of acsr
drake, in Proc. IEEE Winter Power Meeting, New York, Jan. 31Feb.
5, 1971, pp. 18.
[8] R. Claren and G. Diana, Dynamics strain distribution on loaded
stranded cables, IEEE Trans. Power App. Syst., vol. PAS-88, no. 11,
pp. 16781690, Nov. 1969.
[9] P. B. McGill and G. E. Ramey, Effect of suspension clamp geometry
on transmission line fatigue, J. Ener Eng., ASCE, vol. 112, no. 3, pp.
168184, 1986.
[10] G. E. Ramey, Conductor Fatigue life research eolian vibration of transmission lines, Jan. 1987, EL-4744, EPRI Res. Project 1278-1, Final
Rep.
[11] T. Sepp, Transmission Line Vibration IV, Fatigue Tests of ACSR Ibis
Conductor. Helsinki, Finland: Imatran Voima Osakeyhtio, 1968.
[12] A. R. Hard, Studies of conductor vibration in laboratory span, outdoor
test span and actual transmission lines, CIGRE, Rep. 404, 1958.
[13] J. C. Poffenberger and R. L. Swart, Differential displacement and dynamic conductor strain, IEEE Trans. Power App. Syst., vol. PAS-84,
no. 4, pp. 281289, Apr. 1965.
[14] IEEE comittee report, Standardization of conductor vibration measurements, IEEE Trans. Power App. Syst., vol. PAS-85, no. 1, pp.
1022, Jan. 1966.
[15] R. A. Komenda and R. L. Swart, Interpretation of field vibration data,
IEEE Trans. Power App. Syst., vol. PAS-87, no. 4, pp. 10661073, Apr.
1968.
[16] R. H. Scanlan and R. L. Swart, Bending stiffness and strain in stranded
cables, presented at the IEEE Winter Power Meeting, New York, 1967.
Sylvain Goudreau received the Bachelor and Master degrees in mechanical
engineering from cole Polytechnique de Montral, Montral, QC, Canada, in
1977 and 1980, respectively, and the Ph.D. degree in mechanical engineering
from Universit Laval in 1990.
He is a Professor at the Department of Mechanical Engineering at Universit Laval, Qubec City, QC, Canada, and Principal Researcher of Groupe de
REcherche en Mcanique des Conducteurs Ariens Research Group. He is a
Registered Professional Engineer in the Province of Qubec. Before beginning
his Ph.D. studies, he was with the National Research Council of Canada (Institut
du gnie des matriaux, Montral), where he was involved in the development
of a mechanical testing laboratory and in studies on composite materials. In
1988, he joined the Mechanical Engineering Department at Universit Laval.
His research activities are in the field of mechanical behavior of overhead line
conductors and their related fatigue problems. He is author or coauthor of many
technical reports and papers on these subjects.

IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 25, NO. 4, OCTOBER 2010

Frdric Lvesque is a Postdoctoral Fellow at the Department of Civil Engineering, Sherbrooke University, Sherbrooke, QC, Canada. He received the
B.Sc., M.Sc., and Ph.D. degrees in mechanical engineering from the University Laval, Qubec City, QC, Canada .
His research interests are the mechanics involved in the fatigue of overhead
electrical conductors (contact and fracture mechanics, stress analysis, modelization, and mitigation of aeolian vibrations).

Alain Cardou received the M.Sc. and Ph.D. degrees in mechanics and materials
from the University of Minnesota at Minneapolis.
He is Adjunct Professor and, formerly, Head of the Department of Mechanical
Engineering at Universit Laval, Qubec City, QC, Canada. His general research
interests are stress and strength analysis, of which he is the author or coauthor
of many papers. For several years, in collaboration with some power utilities
and within the GREMCA research group, he has been working on overhead
electrical conductor fatigue problems.
Dr. Cardou is a Registered Professional Engineer in the Province of Qubec,
a Fellow of the Canadian Society for Mechanical Engineering, and a member
of the American Academy of Mechanics.

Louis Cloutier received the Ph.D. degree in mechanical engineering from


Universit Laval, Qubec City, QC, Canada, in 1966, and pursued postdoctoral
work in contact mechanics for a year at Cambridge University, Cambridge,
U.K.
He held different functions in research laboratories (National Research
Council of Canada and IREQ [Hydro-Qubecs Research Institute]), industries
(Gleason Works, Roctest Ltd., and Sogequa, Inc.), and universities (Laval and
Sherbrooke). His professional experience of more than 40 years led him to
work in several projects related to mechanical power transmission, medical
instrumentation, and electrical power transmission. In the last field, his interests
have been mainly devoted to problems related to transmission-line mechanics:
conductors, insulators, spacers, accessories, and more recently, line supports.
He is the author or coauthor of several publications in related fields and holds
two patents. He was Chair Holder from 2004 until his retirement in 2009 of the
industrial chair created by Hydro Qubec Transnergie in collaboration with
the Natural Sciences and Engineering Research Council of Canada for studies
of the structural and mechanical aspects of overhead transmission lines. He is
Adjunct Professor at the Department of Mechanical Engineering of Universit
Laval.
Dr. Cloutier is a member of lOrdre des Ingnieurs du Qubec, an active
member of several technical and learned societies, a Distinguished Member of
CIGRE, and Fellow of the Canadian Society for Mechanical Engineering.

You might also like