Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Failure of Flawed Utility Poles in

Wind Gusts
The purpose of this study is to account for failures of wood utility poles in wind storms
based on dynamic analysis and pole imperfections. The utility pole supporting multiple
James F. Wilson overhead transmission lines is modeled as a uniform Bernoulli–Euler cantilevered beam
PE fixed at the base and subjected to three types of suddenly applied transverse loads that
Professor Emeritus simulate a wind gust: a uniform pole pressure, a point load at the tip accounting for line
Department of Civil and Environmental and transformer drag, and another point load near midlength, accounting for drag on
Engineering, lines strung from that location. The dynamic pole moments are based on normal mode
Duke University, calculations rather than static calculations with a dynamic impact factor, and the critical
6319 Mimosa Drive, flexural stresses include stress concentrations arising from pole imperfections such as
Chapel Hill, NC 27514 holes, knots, and surface gouges. A case study illustrates the results for one of about 400
e-mail: jwilson@duke.edu failed wood poles downed in a single New Jersey storm in 2003 with 107 km/h (67 m/h)
wind gusts. Here, the critical pole stress based on the dynamic model and a hole imper-
fection exceeded the proportional limit stress of the wood. The predicted dynamic stresses
are higher than those based on the National Electrical Safety Code.
关DOI: 10.1115/1.4002763兴

Keywords: beam dynamics, normal mode analysis, stress concentrations, utility poles,
wind loading, wood pole failure

1 Introduction The general intent is to use this study to check similar pole de-
signs based on the design codes alone. In this way, perhaps the
In countries with ample forests, chemically treated wood utility
2006 western Kansas disaster could have been avoided, an event
poles continue to be the most economical structure to support
in which a 196 km/h 共122 m/h兲 wind gust accounted for ground-
overhead electric distribution lines, telephone wires, and TV
line breaks of the 21 m 共69 ft兲 tall Douglas fur poles along a 3.2
cables. Properly designed and periodically inspected 共generally
km 共2 mi兲 straight stretch 关9兴. Just one pole, weakened with a
once every 10 years兲, a 15.2 m 共50 ft兲 pole with a 0.305 m 共1 ft兲
ground-line stress concentration, could have precipitated this
mean diameter may have a lifetime of 60 or more years. As dis- domino-type event.
cussed by Pellicane and Franco 关1兴, who include about 20 histori-
cal citations on the subject, the load-carrying capacity of wood
utility poles for a given transverse loading is mainly affected by
ground-induced decay and loose knots and holes, whereas grain 2 Mathematical Model
orientation, taper, and moisture content have minor effect on pole An example of a cantilevered wood utility pole unsupported by
material failure. These factors are tempered by the species of guy wires and subjected to a wind gust perpendicular to the over-
wood 共Eastern cedar, Douglas fir, southern pine, etc.兲 and their head distribution lines is shown in Fig. 1. A planar model of this
mechanical properties. Sometimes radial holes are bored near the pole, its coordinate system, and its component wind gust loads P,
ground-line holes created as pathways to induce chemical preser- Q, and q0, are depicted in Fig. 2共a兲. The tip point load P accounts
vatives, which slow decay. Elkins et al. 关2兴 suggested patterns of for the wind drag on both the transformer and on the distribution
such through-boring in experiments using 1.27 cm holes and de- lines at the tip and facing the wind; the point load Q represents the
duced particular staggered hole patterns near the pole base that do wind drag load on lines facing the wind and located at a distance
not reduce the bending strength by more than 5%. In general, the x = b from the base; and q0 represents a uniform transverse load
designer of wood utility poles relies on data from several sources: over the whole exposed pole of length ᐉ. These three loads are
for the mechanical properties, the American National Standards modeled as step loads as shown in Fig. 2共b兲, conservatively as-
Institute 共ANSI兲 关3兴 and the U.S. Department of Agriculture sumed to be in phase and applied at time t = 0. That is P共t兲
关4–6兴; for estimates of local increases in bending stress 共stress
= Q共t兲 = q共t兲 = 0 for t ⬍ 0 and
concentrations兲 near holes and imperfections such as surface
gouges, the results summarized by Pilkey and Pilkey 关7兴; for es- P共t兲 = P0 ; Q共t兲 = Q0 ; q共t兲 = q0 for t⬎0 共1兲
timates of extreme wind speeds occurring throughout the United
The pole is assumed to be uniform and linear elastic with a bend-
States, the maps of the National Electrical Safety Code 关8兴.
ing stiffness EI and a transverse dynamic displacement coordinate
The present study addresses the dynamic analysis of a uniform
of v = v共x , t兲. The transverse motion of the pole is represented by
cantilevered pole without guy wires, subjected to suddenly ap-
plied uniform and point transverse loads simulating a wind gust. the Bernoulli–Euler beam equation, as derived from classical
The critical dynamic bending stress, calculated by solving the beam theory 关10兴. That is,
partial differential equation for the transverse response of the pole, ⳵ 4v ⳵ 2v
is modified with possible stress risers due to holes and surface EI 4 + m̄ 2 = q̄共x,t兲 共2兲
⳵x ⳵t
groves. A case study is then used to illustrate the theory of failure.
Here, m̄ is the mass per unit length of the pole and may include
the mass of the pole attachments such as a transformer and hang-
Contributed by the Reliability, Stress Analysis, and Failure Prevention Committee ing wires uniformly distributed along the pole length. The loading
of ASME for publication in the JOURNAL OF MECHANICAL DESIGN. Manuscript received
December 31, 2009; final manuscript received July 30, 2010; published online q̄共x , t兲 includes all three types of wind loads. At the fixed base, the
November 12, 2010. Assoc. Editor: Steven J. Skerlos. respective end conditions for zero displacement and zero slope are

Journal of Mechanical Design Copyright © 2010 by ASME NOVEMBER 2010, Vol. 132 / 111006-1

Downloaded From: http://mechanicaldesign.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


XIV − ␣4X = 0 共6兲
where the pole frequency ␻, written in terms of the parameter ␣ᐉ
is given by

␻ = 共␣ᐉ兲2 冑 EI
m̄ᐉ4
共7兲

The boundary conditions of Eqs. 共3兲 and 共4兲 as they apply to Eq.
共6兲 is now reduce to

X共0兲 = X⬘共0兲 = 0; X⬙共ᐉ兲 = X⵮共ᐉ兲 = 0 共8兲


The four independent solutions of Eq. 共6兲 are given by

X = D1 sin ␣x + D2 cos ␣x + D3 sinh ␣x + D4 cosh ␣x 共9兲


in which D1, D2, D3, and D4 are constants. When each condition
共8兲 is applied to Eq. 共9兲, four homogeneous algebraic equations
give a nonzero solution for X only for discrete roots ␣nᐉ, which
Fig. 1 Utility poles with a wind gust perpendicular to the over- satisfy
head lines
cos ␣nᐉ cosh ␣nᐉ = − 1 共10兲
The roots 共␣nᐉ兲 of Eq. 共10兲 are given by
⳵ v共0,t兲 ␣1ᐉ = 1.875, ␣2ᐉ = 4.694, ␣3ᐉ = 7.855, ␣nl = ␲共2n
v共0,t兲 = 0, =0 共3兲
⳵ x2 − 1兲/2, n = 3,4,5, . . . 共11兲
The respective conditions for zero moment and zero transverse
shear at the tip are With Eq. 共11兲, the n frequencies ␻ = ␻n can be calculated from Eq.
共7兲 and its corresponding free vibration normal modes are solu-
⳵2v共ᐉ,t兲 ⳵3v共ᐉ,t兲 tions 共9兲 that satisfy Eq. 共8兲, or
EI = 0, EI =0 共4兲
⳵ x2 ⳵ x3
cos ␣nᐉ + cosh ␣nᐉ
The solution of Eq. 共2兲 for the stated loadings and end condi- Xn共x兲 = cos ␣nx − cosh ␣nx − 共sin ␣nx
sin ␣nᐉ + sinh ␣nᐉ
tions is derived using the normal mode method, which is well
documented 关10,11兴. First, the discrete free vibration frequencies − sinh ␣nx兲 共12兲
␻ = ␻n and the accompanying modal displacements X = Xn共x兲 are Return now to the problem of wind gust loading. Assume a
computed. Thus, assume that transverse harmonic motion has the solution to Eq. 共4兲 as a product of the normal modes 共12兲 and a
form generalized coordinate y n 共t兲 yet to be determined. That is, let
v = X sin ␻t 共5兲 ⬁
When Eqs. 共5兲 and 共2兲 are combined, the result is v共x,t兲 = 兺 X 共x兲y 共t兲
n=1
n n 共13兲

When Eq. 共13兲 is combined with Eq. 共2兲 and the conditions for
orthogonality of Xn共x兲 are applied, then the result is

d2y n共t兲
+ ␻n2y n共t兲 = pn共t兲 共14兲
dt2
where

冕 冕
ᐉ ᐉ
1
pn共t兲 = Xn共x兲q̄共x,t兲dx; Cn = X2n共x兲dx 共15兲
m̄Cn 0 0

If the pole is initially at rest just before the wind gust, then
v共x , 0兲 = ⳵v共x , 0兲 / ⳵t = 0. These two initial conditions imply that
y n共0兲 = dy n共0兲 / dt = 0 so that the solution to Eq. 共14兲 takes the fol-
lowing form of the Duhamel integral


t
1
y n共t兲 = pn共␶兲sin ␻n共t − ␶兲d␶ 共16兲
␻n 0

When the three components of q̄共x , t兲 defined by Eq. 共1兲 are


combined with Eq. 共15兲, pn共t兲 is computed

pn共t兲 =
1
m̄Cn 冋 P0Xn共ᐉ兲 + Q0Xn共b兲 + q0 冕

0
Xn共x兲dx 册 共17兲

Fig. 2 „a… Planar model of the failed utility pole 60103 and „b… When Eq. 共17兲 is combined with Eq. 共16兲 and integrations are
step approximations to the pole wind gust loads performed, the generalized coordinate becomes

111006-2 / Vol. 132, NOVEMBER 2010 Transactions of the ASME

Downloaded From: http://mechanicaldesign.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


冕 冉 冊 冕冉 冊 冕冉 冊
ᐉ b ᐉ
P0 Q0 1 ⳵ 2v 2
1 ⳵v 2
1 ⳵v 2
y n共x兲 = Xn共ᐉ兲共1 − cos ␻nt兲 + Xn共b兲共1 − cos ␻nt兲 V= EI dx − mbg dx − mtg dx
m̄␻n Cn 2
m̄␻n Cn2
2 ⳵ x2 2 ⳵x 2 ⳵x


0 0 0

q0 共26兲
+ 共1 − cos ␻nt兲 Xn共x兲dx 共18兲
m̄␻n2Cn 0 The terms on the right in Eq. 共26兲 represent, respectively, the
pole’s bending energy, the loss in potential energy due to the
These results are summarized. The chosen number n of pole
vertical drop of mbg of the wires attached at x = b, and the coun-
free vibration frequencies are computed from Eq. 共7兲 for the con-
terpart energy loss for mtg at x = ᐉ. The maximum values of the
secutive roots given by Eq. 共11兲. The pole’s transverse displace-
integrands in Eq. 共26兲, computed from Eq. 共22兲, are
ment distribution v共x , t兲, starting at time zero when the wind gust
begins, is computed from Eq. 共13兲 using the normal modes of Eq.
共12兲 and the generalized coordinate solution of Eq. 共18兲. Further, 冉 冊
⳵v
⳵x MAX
= v0

2ᐉ
sin
␲x
2ᐉ
; 冉 冊 ⳵ 2v
⳵ x2 MAX
= v0 冉 冊

2ᐉ
2
cos
␲x
2ᐉ
the pole’s transverse bending moment distribution M共x , t兲 is com-
puted using Eq. 共18兲 and the second derivative of Eq. 共12兲. That 共27兲
is, With Eq. 共27兲, the maximum potential energy is computed from
n n Eq. 共26兲 as

冉 冊
⳵2Xn共x兲
M共x,t兲 = EI 兺 ⳵ x2 兺
y n共t兲 = EI Xn⬙共x兲y n共t兲 共19兲
VMAX =
␲4
3 − m bg
␲2 b ᐉ
− sin
␲b
− m tg
␲2
共28兲
n=1 n=1
32ᐉ 4ᐉ 2 2␲
2
2ᐉ 8ᐉ
in which According to the Rayleigh method, ␻0 is determined by equating
Xn⬙共x兲 = ␣n 2
冋 − cos ␣nx − cosh ␣nx −
cos ␣nᐉ + cosh ␣nᐉ
sin ␣nᐉ + sinh ␣nᐉ
共− sin ␣nx
the maximum kinetic energy 共25兲 to the maximum potential en-
ergy 共28兲. The result is as follows:

− sinh ␣nx兲 册 共20兲 8ᐉ 4ᐉ



␲2 ␲2EI
2 − m bg
b 1 ␲b
− sin
ᐉ ␲ 2ᐉ
冉 − m tg 冊 册
冉 冊 冉 冊
␻ 02 = 共29兲
3 4 ␲b 2
m̄ᐉ − + mb 1 − cos + mt
2 ␲ 2ᐉ
3 Alternative Pole Frequency Analysis
Consider the Rayleigh frequency for the particular case of a
The above modal analysis that leads to the pole bending fre- uniform pole without attachments or mb = mt = 0. Then Eq. 共29兲
quencies given by Eqs. 共7兲 and 共11兲 does not include the effects of reduces to


several direct longitudinal pole compressive loads: the combined
load mtg of the transformer and its wires at the tip and the load EI
mbg of the hanging wires at x = b. How these compressive load- ␻0 = 3.6638 共30兲
m̄ᐉ4
ings affect the fundamental bending frequency ␻1 predicted by
Eqs. 共7兲 and 共11兲 is now investigated using the Rayleigh method. This frequency is 4.2% higher than the “exact” fundamental fre-
Numerous examples of the Rayleigh method, which predict the quency given by Eqs. 共7兲 and 共11兲 or
frequencies ␻0 as an upper bounds for ␻1, are illustrated by Wil-
son 关10兴 in Chap. 2 in the context of offshore pilings.
In the Rayleigh method, one chooses a simple mode shape ␺共x兲
␻1 = 3.5156 冑 EI
m̄ᐉ4
共31兲

of amplitude v0 that satisfies the geometric boundary conditions of Further, Eq. 共29兲 shows that the fundamental frequency decreases
deflection and slope. For the cantilevered pole, such a shape is as the discrete loads at x = b and/or at x = ᐉ increase. In the limiting


␺共x兲 = v0 1 − cos
␲x
冊 共21兲
case, ␻0 = 0 where the numerator of Eq. 共29兲 is zero and the pole
buckles under the critical load given by

冉 冊
2ᐉ
␲2EI b 1 ␲b
which satisfies the base fixity ␺共0兲 = ␺⬘共0兲 = 0 and also the tip Pcr = = mbg − sin + m tg 共32兲
conditions ␺共ᐉ兲 ⫽ 0 , ␺⬘共ᐉ兲 ⫽ 0. Using Eq. 共21兲, harmonic motion 4ᐉ2 ᐉ ␲ 2ᐉ
in the first frequency is characterized by It is observed that ␲2EI / 共4ᐉ2兲 is precisely the well-known Euler


v共x兲 = v0 1 − cos
␲x
2ᐉ
冊sin ␻0t 共22兲
column buckling load for an end-loaded cantilevered beam, de-
rived from statics. See Timoshenko and Gere 关12兴.
These results complement the modal analysis. In the case study
The kinetic energy K for the pole based on the separate mass that follows, it will be shown that the discrete loads at x = b and
components and their respective velocities ⳵v / ⳵t is x = ᐉ are well below the Euler buckling load, adding to the cred-

冕 冉 冊 冉 冊 冉 冊

ibility that the modal analysis gives reasonably accurate dynamic
1 ⳵v 2
1 ⳵v 2
1 ⳵v 2
pole moment and bending stress responses if longitudinal effects
K= m̄ dx + mb + mt 共23兲 of the transformer and wire loads are ignored.
2 0
⳵t 2 ⳵t x=b 2 ⳵t x=ᐉ

From Eq. 共22兲, the maximum velocity at location x becomes 4 Typical Pole Flaws and Critical Stresses

冉 冊 ⳵v
⳵t MAX

= v0␻0 1 − cos
␲x
2ᐉ
冊 共24兲
The dynamic flexural stress at the outer surface of a solid, uni-
form pole of diameter D can be computed from classical beam
theory and the moment given by Eq. 共19兲, with Eq. 共20兲. That is,
With Eq. 共24兲 the maximum kinetic energy is computed from Eq.
32M共x,t兲
共23兲 as ␴共x,t兲 = Ktg 共33兲

冋 冉 冊 冉 冊 册
␲D3
3 4 ␲b 2
KMAX = ␻02 m̄ᐉ − + mb 1 − cos + mt 共25兲 Following the nomenclature of Pilkey and Pilkey 关7兴, Ktg is a
2 ␲ 2ᐉ constant that defines a stress riser or stress concentration factor
The potential energy V of the pole, together with that of its that accounts for an increase in the nominal flexural stress when
transformer and hanging wires, is there is a flaw such as a hole or gouge in the pole. Note that the

Journal of Mechanical Design NOVEMBER 2010, Vol. 132 / 111006-3

Downloaded From: http://mechanicaldesign.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Fig. 4 Stress concentration at the corner A of a shallow notch
or gouge in local section of an isotropic circular cylinder. Ap-
plied to a utility pole, the bending moment M is caused by the
wind gust, which is perpendicular to the pole in the left view
Fig. 3 Stress concentration factor at a through-hole in a local „data adapted from Pilkey and Pilkey †7‡….
section of an isotropic circular cylinder. Applied to a utility
pole, the bending moment M is caused by the wind gust, which
is perpendicular to the pole in the left insert „data adapted from Shown in Fig. 5共a兲 is part of the left rear of the vehicle at rest
Pilkey and Pilkey †7‡…. in the oncoming lane, together with the downed lines ahead of the
vehicle, and a cut portion of the ruptured pole. The new standing
pole that replaced the broken one is to the left of the vehicle. Both
absence of such flaws implies that Ktg = 1. For a pole of diameter poles were southern yellow pine treated with preservative, with a
D with a transverse through-hole of diameter d, data for Ktg as a nominal diameter D = 0.305 m 共12 in.兲, an above-ground height of
function of the ratio d / D were compiled by Pilkey and Pilkey 关7兴,
and selected data are shown in Fig. 3. Here, it is observed that
Ktg → 3.0 as d / D → 0 and Ktg → 4 as d / D → 0.3, a practical upper
limit for present applications. Another type of flaw is shown in
Fig. 4: a longitudinal gouge of width D / 4 and a depth of D / 8. If
D ⬎ 0.165 m 共0.54 ft兲, as for common utility poles, then Ktg
= 2.1 at the gouge fillet.
Strictly speaking, the data of Figs. 3 and 4 are applicable only
to isotropic cylinders, whereas wood poles are orthotropic and
strongest in the longitudinal direction. However, Ktg values have
yet to be measured for flawed wood poles in flexure. Thus, it is
judicious to use these referenced stress concentration factors as
approximations rather than to completely ignore such flaws in the
pole design process.

5 Case Study
On the afternoon of November 13, 2003, a 1999 Chevrolet Trail
Blazer was traveling in the northbound lane of South Bridge
Street in Somerville, NJ. As the vehicle approached the utility
pole no. 60103, which had been in service for 10 years, and lo-
cated on the west edge of the road, the pole broke in two at c
= 4.6 m 共15 ft兲 above the ground. The top section of this pole with
its attached wires and transformer fell across South Bridge Street
and onto the hood and windshield of the oncoming Trail Blazer.
The forward momentum of the struck Trail Blazer as it pulled the
entangled wires, together with a high wind, led to the toppling of
the five consecutive power poles on the west edge of South Bridge
Street behind the Trail Blazer. Weather records for Somerville, NJ
at the time of this accident indicated that the temperature was
44 ° F and that a wind of speed V = 107 km/ h 共67 m/h兲 swept Fig. 5 „a… In the foreground is a section cut from the ruptured
across an open park at a right angle to the line of the South Bridge pole 60103. „b… The ruptured pole section shows longitudinal
Street utility poles. shear failure.

111006-4 / Vol. 132, NOVEMBER 2010 Transactions of the ASME

Downloaded From: http://mechanicaldesign.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Table 1 Computed loads and dynamic responses for pole No. 60103 to a 107 km/h „67 m/h…
wind gust of 2 s duration

Type of load or stress SI units English units

Wind pressure on pole 1.154 kPa 24.1 lb/ ft2


q0: wind load per unit pole height 351 N/m 24.1 lb/ft
P0: wind load on wires and transformer at top of pole 1.41 kN 316 lb
Q0: wind load on wire at b = 5.18 m 共17 ft兲 above base 3.73 kN 839 lb
Lowest pole bending frequency 共n = 1兲 0.903 Hz 0.903 Hz

Pole’s maximum horizontal displacements


At top 1.30 m 4.28 ft
At c = 4.6 m 共15 ft兲 above base 0.234 m 0.767 ft

Maximum flexural stress


At base 共no stress concentration兲 55.3 MPa 8,017 psi
At c = 4.6 m 共15 ft兲, with no stress concentration 26.8 MPa 3,890 psi
At c = 4.6 m 共15 ft兲, incl. hole stress concentration 73.7 MPa 10,700 psi
Proportional limit, southern yellow pine 关4兴 53.8 MPa 7,800 psi
Rupture stress, southern yellow pine 关4兴 88.4 MPa 12,820 psi

13.7 m 共45 ft兲, and a rupture strength of 88.4 MPa 共12,820 psi兲 pole diameter D in the range of 0.05ⱕ d / D ⱕ 0.1, then the stress
关4兴. The poles, located in an urban setting, were packed in solid concentration factor from Fig. 2 is Ktg ⬇ 2.75 and the peak tensile
ground between a concrete sidewalk and the road’s concrete curb, flexural stress at the edge of the hole on the wind side is ␴
which justifies the clamped base condition assumed in the above = Ktg共26.8兲 = 73.7 MPa 共or 10,700 psi兲, a value that exceeds the
dynamic analysis. Shown in Fig. 5共b兲 is a close up view of the pole’s proportional limit but is just under its rupture strength. It is
ruptured pole, which is a typical longitudinal shear failure. noted that a loose knot near the rupture point could have had the
The magnitudes of the three classes of wind drag loads same effect as this hole.
共q0 , P0 , Q0兲 are based on the reported wind speed of 107 km/h, Consider now how the ruptured pole measures up to the design
which is assumed to be the peak gust velocity. The customary specifications of the National Electrical Safety Code 共the code兲.
empirical expression for the total wind drag force FD at right As shown in Fig. 250-2 of the code, the design wind speed is 129
angles to a cylindrical object, which is stationary, or nearly so, is km/h 共80 m/h兲 for central New Jersey, the region of this case
given by study. With this wind speed and the formula given in Section 25,
1 paragraph 250-C of the code, the wind pressure was computed as
F D = C D 2 ␳ V 2A P 共34兲 784 Pa 共16.38 lb/ ft2兲. Further, the code specifies an overload fac-
Here, CD = 2.0 is the drag coefficient for cylinder, which is long tor for q0 of 4.0 and an overload factor for the transformer and for
compared with its diameter 共the present case兲, ␳ = 1.324 kg/ m3 is all wires of 2.0.
the mass density of the air, and A p = Dᐉ is the projected area of the Using the wind-generated pressure together with the geometry
pole perpendicular to the wind gust. With numerical values of A P, of the pole with its attachments, the loads q0, P0, and Q0, includ-
for all wires extending to the midlength of its two adjacent poles ing their appropriate overload factors, were computed and are
and for the transformer at the top of the pole, the three respective listed in Table 2. For a flawless, uniform pole fixed at its base, the
drag loads q0 , P0 , Q0 were calculated using Eq. 共34兲. These loads maximum flexural stress always occurs at the base. This calcu-
were assumed to be in phase because the wind gust was most lated stress for the pole is 52.5 MPa 共7620 psi兲, which is listed
likely uniform as it swept unobstructed across the open park ad- near the end of Table 2. Also listed is the flexural stress for rupture
jacent to this line of utility poles of South Bridge Street. of southern yellow pine, which is 88.4 MPa 共12,820 psi兲. Since
The results based on the dynamic analysis are summarized in the computed base stress is less than both the pole’s proportional
Table 1, in dual units. These were based on the pole’s Young’s limit and its rupture strength, pole no. 60103 met the code require-
modulus of E = 12.4 MPa and its equivalent mass per unit length ments when it was installed in 1993.
of m̄ = 62.2 kg/ m. This equivalent unit mass was the sum of 37.2
kg/m for the pole alone and 19.8 kg/m, where the latter was the
transformer mass with that of its surrounding wires, modeled as a
uniform load intensity along the pole length.
The horizontal displacement and bending moment responses 6 Discussion
along the pole as a function of time for up to 2 s after the wind This study of wood utility poles illustrates the role of wind
gust loading were computed utilizing a program written in the gusts, pole dynamics, and pole flaws 共bored holes, knots, and
object-oriented language of MATHEMATICA 关13兴. Suitable series gouges兲 on the pole’s structural integrity. Five important features
convergence was achieved using the first three modes 共n and limitations of this study are now discussed.
= 1 , 2 , 3兲. Listed in Table 1 are the lowest pole bending frequency First, consider wind gusts. Measurements show that short wind
共␻1 = 5.67 rad/ s = 0.903 Hz兲, the maximum dynamic flexural gust velocities of 1–5 s in duration are usually 20–30% higher
stresses at the base 共55.3 MPa兲 and at the rupture point 共26.8 than the measured mean wind velocity 关14兴. Thus, if for design
MPa兲, and stresses that are below the pole’s rupture stress of 88.4 purposes the expected extreme horizontal wind speed is based on
MPa. These calculations lead to the important conclusion that, had the maps presented in the codes that wind speed should be en-
pole no. 60103 been free of all loose knots and through-holes near hanced by at least a factor of 1.2. For the case study, for instance,
the break, the pole would have maintained its integrity and not it is not clear whether or not the 107 km/h 共67 m/h兲 wind speed
snapped in the reported wind gust of 107 km/h 共67 m/h兲. reported by the National Weather Service included the gust factor.
However, consider a through-hole in the neighborhood of the If not, then the short-time wind speed when pole 60103 ruptured
4.6 m 共15 ft兲 point of the rupture, a hole used to bolt the lower was actually about V = 共1.2兲 共107兲 = 128 km/ h 共80 m/h兲. Since the
four cables to the pole. For the ratio of bolt hole diameter d to pole flexural stresses vary as the square of the wind speed, then

Journal of Mechanical Design NOVEMBER 2010, Vol. 132 / 111006-5

Downloaded From: http://mechanicaldesign.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Table 2 Design loads and stresses for pole no. 60103 based on the code †8‡. The design wind
speed was 129 km/h „80 m/h….

English Code or
Type of load or stress SI units units reference

Wind pressure on pole 784 Pa 16.38 lb/ ft2 Table 250-2


q0: wind load per unit pole length 共includes 4.0 overload factor兲 956 N/m 65.52 lb/ft Table 261-3A
P0: wind load on wires and transformer at top of pole
共includes 2.0 overload factor兲 10.3 kN 2,325 lb Table 261-3A
Q0: wind load on wires at b = 5.18 m 共17 ft兲
共includes 2.0 overload factor兲 11.7 kN 2,635 lb Table 361-3A
Maximum flexural stress at base 52.5 MPa 7,620 psi 共Computed兲
Proportional limit, southern yellow pine 53.8 MPa 7,800 psi 关4兴
Rupture stress for southern yellow pine 88.4 MPa 12,820 psi 关4兴

the pole’s flexural stress at the rupture point c = 4.6 m 共15 ft兲 11, respectively. With this equivalent pole length, the full clamped
would be 共1.2兲2共73.7兲 = 106 MPa 共15,410 psi兲, which exceeds the base occurs beneath the soil surface where this depth depends on
rupture strength of the pole as listed in Table 1. the strength properties of the soil. Use of the present dynamic
Second, consider the importance of a dynamic analysis. This analysis is justified if such an equivalent pole length is used for
general analysis applied to the case study led to a pole natural poles buried in nonfirm soil.
frequency of 0.903 Hz. Thus, the pole’s natural period was 1.11 s, Fifth, consider pole flaws. Sometimes through-holes need to be
which falls in the range of 0.5–5 s for typical wind gusts. Such a tolerated, either below the base for the injection of wood preser-
matching of periods could lead to a pole resonance state with vatives or at the bolted brackets needed to hold the distribution
enhanced responses. For instance, from Table 1, the peak flexural wires. Sometimes surface gouges resulting from installation or
stress at the base was calculated as 55.3 MPa, which exceeds its from vehicle impact need to be accommodated. In such cases, the
counterpart stress of 52.5 MPa 共Table 2兲 using the code with its pole’s flexural stress calculated using the bending moment of a
several somewhat arbitrary overload factors. The dynamic analy- dynamic analysis is multiplied by the appropriate stress concen-
sis can be considered a calculation based purely on fundamental tration factor of Fig. 3 or by Ktg = 2.1 for a gouge approximated by
mechanics without any overload factors. the geometry of Fig. 4. These referenced values of Ktg are strictly
Third, consider the pole geometry assumed for the dynamic applicable to isotropic materials, but lacking further data, are con-
model. In the case study, the pole diameter decreased by about sidered as reasonably realistic when applied to flawed orthotropic
20% from base to tip, which is typical for utility poles in use. The wood utility poles.
pole diameter D was chosen herein as the mean value for the In summary, wood utility pole design should always include
gradually tapered pole. In this way, the dynamic results could be considerations of wind gust loading, pole dynamics, and pole
compared with those static responses derived from the code in flaws acting as stress risers. The uniform pole of mean diameter to
which a uniform pole of mean diameter is also assumed. However, represent a gradually tapered pole is appropriate for analysis and
in calculating the flexural stresses based on the dynamic moment, design purposes. For future sustainable designs, the resulting criti-
the local diameter of the tapered pole was used. cal stresses can be used to complement the code-based designs.
Further calculations led to bounds for the bending frequencies
of a gradually tapered pole: upper bounds for a uniform pole of Acknowledgment
diameter 1.1D and lower bounds for a uniform pole of diameter The author thanks Hugh M. Turk for supplying the evidence
0.9D. In the first case, these n frequencies of Eqs. 共7兲 and 共11兲 are related to the case study and for the partial financial support dur-
modified as ing this accident investigation.

␻n = 共␣nᐉ兲2 冑 共1.1兲4EI
共1.1兲2m̄ᐉ4
= 1.1共␣nᐉ兲2 冑 EI
m̄ᐉ4
共35兲 Nomenclature
A p ⫽ projected area of wire cluster or pole in direc-
which are 10% higher than those based on the mean diameter D. tion of wind velocity
In like manner, the lower bound frequencies for the minimum pole b ⫽ height of lowest wire cluster above ground
diameter are calculated to be 10% lower than those based on D. CD ⫽ coefficient of wind drag on a long, solid
For practical purposes, these results strongly suggest that the cylinder
mean pole diameter is an adequate physical representation of a c ⫽ height of pole rupture point above ground
tapered pole whose respective frequencies lie between those of the D ⫽ pole diameter
larger and smaller poles or close to the frequencies based on the EI ⫽ pole flexural stiffness
mean diameter. Also, the fundamental periods for these poles, FD ⫽ total drag force on a cylindrical solid
large, mean, and small, all lie within the range of typical gust Ktg ⫽ stress concentration factor, Figs. 3 and 4
periods: 0.5–5 s. A future dynamic analysis of tapered poles would ᐉ ⫽ pole length above ground
involve an extensive numerical analysis with nonorthogonal mode m̄ ⫽ pole mass per unit length
shapes rather than the closed form solutions herein. A starting mbg ⫽ weight of pole wires at x = b
point for such an analysis would be the analytical results dis-
mtg ⫽ weight of transformer and wires at x = ᐉ
cussed by Wilson 关15兴 concerning the toppling of truncated cones
and trees. n ⫽ mode number 共n = 1 , 2 , 3 , . . .兲
Fourth, consider the pole’s base fixity. The fully clamped con- P = P共t兲 ⫽ horizontal wind drag load at tip of pole
dition was applicable to the case study because this pole was p共t兲 ⫽ modal load
buried in firm soil between a concrete walk and curb. For cases in Q = Q共t兲 ⫽ horizontal wire cluster drag load at x = c
which the soil base is softer 共in sand for instance兲, an equivalent q0 ⫽ uniform component of pole drag load per unit
longer length for the pole can be chosen using either the method length
of Kocsis or Reese, as presented by Wilson 关10兴 in Chaps. 2 and

111006-6 / Vol. 132, NOVEMBER 2010 Transactions of the ASME

Downloaded From: http://mechanicaldesign.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


q̄共x , t兲 ⫽ total pole loading per unit length 关4兴 USDA, 1954, Wood Handbook 72, Forest Products Laboratory, USDA, Wash-
ington, DC.
Xn = Xn共x兲 ⫽ nth mode shape of pole 关5兴 Koch, P., 1985, Agricultural Handbook 605, USDA, Washington, DC.
x ⫽ pole length coordinate 关6兴 Wolfe, R. W., Bodig, J., and Lebow, P. K., 2001, “Derivation of Nominal
V ⫽ wind speed Strength for Wood Utility Poles,” Forest Service, USDA, Technical Report No.
FPL-GTR-128.
v = v共x , t兲 ⫽ transverse pole displacement 关7兴 Pilkey, W. D., and Pilkey, D. F., 2008, Peterson’s Stress Concentration Fac-
y n共t兲 ⫽ nth generalized coordinate of pole motion tors, 3rd ed., Wiley, Hoboken, NJ.
关8兴 IEEE, 1993, National Electrical Safety Code, IEEE, New York, pp. 143–153.
␣n ⫽ nth frequency parameter 关9兴 Engle, M., Brown, R., Phillips, E., and Bingel, N., 2009, “Extreme Winds Test
␦共t兲 ⫽ delta function Wood Pole Strength,” http://www.tdworld,com/overhead_transmission/
␳ ⫽ air mass density power_extreme _winds_test/index.html
␻n 关10兴 Wilson, J. F., 2003, Dynamics of Offshore Structures, 2nd ed., Wiley, Hoboken,
⫽ nth pole free vibration frequency, rad/s NJ.
关11兴 Clough, R. W., and Penzien, J., 1993, Dynamics of Structures, 2nd ed.,
McGraw-Hill, New York.
References 关12兴 Timoshenko, S. P., and Gere, J. M., 1961, Theory of Elastic Stability, 2nd ed.,
关1兴 Pellicane, P. J., and Franco, N., 1994, “Modeling Wood Pole Failure,” Wood McGraw-Hill, New York.
Sci. Technol., 28, pp. 261–274. 关13兴 Wolfram, S., 1999, MATHEMATICA Version 4, Wolfram Media, Champaign, IL.
关2兴 Elkins, L., Morrell, J. J., and Leichti, R. J., 2007, “Establishing a Through- 关14兴 Gaythwaite, J., 1981, The Marine Environment and Structural Design, Van
Boring Pattern for Utility Poles,” Wood Fiber Sci., 39共4兲, pp. 639–650. Nostrand Reinhold, New York.
关3兴 American National Standards Institute 共ANSI兲, 2002, American National Stan- 关15兴 Wilson, J. F., 2004, “Segmented Impact Fracture of Toppling Truncated Cones
dard for Wood Poles, ANSI, New York, NY, pp. 1–26. and Tall Trees,” Int. J. Impact Eng., 30, pp. 351–365.

Journal of Mechanical Design NOVEMBER 2010, Vol. 132 / 111006-7

Downloaded From: http://mechanicaldesign.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

You might also like