You are on page 1of 11

Engineering Structures 29 (2007) 821831

www.elsevier.com/locate/engstruct

Numerical and experimental study of vibration mitigation for highway


light poles
Luca Caracoglia a, , Nicholas P. Jones b
a Department of Civil and Environmental Engineering, Northeastern University, Boston, MA 02115, USA
b Whiting School of Engineering, The Johns Hopkins University, Baltimore, MD 21218, USA

Received 16 January 2006; received in revised form 29 June 2006; accepted 29 June 2006
Available online 14 August 2006

Abstract
Highway light poles are slender structures usually characterized by low values of structural damping, a factor that can lead to large-amplitude
vibration sometimes leading to collapse. This paper is motivated by a recent investigation, conducted to identify the reason for repeated failures,
experienced by aluminum tapered light poles in the State of Illinois during a winter storm. The study combined numerical and experimental
full-scale analysis of the structural system and its response to simulated external actions. It was observed that, despite the simple structural form,
the definitive identification of the mechanism causing the oscillations was challenging due to inherent variability in the configuration as well as
the paucity of environmental and response data. However, a plausible mechanism was identified, and a mitigation technique was proposed and
evaluated for amplitude reduction.
c 2006 Elsevier Ltd. All rights reserved.

Keywords: Light poles; Dynamic loads; Vibration; Damping; Wind; Frozen precipitation; Galloping

1. Introduction
Highway light poles are frequently subjected to environmental and wind loads. Despite the simple structural system and
the availability of specific design tools (e.g., [1]), particular occurrences of unpredictable load configurations can lead to the
failure of such systems or their sub-components. This paper is
motivated by a research project managed by the Illinois Department of Transportation, initiated to understand the nature of
some recent failures experienced on light pole structures in both
serviceability and strength (collapse). One of the problems was
the occurrence of large oscillation amplitudes experienced during a wind storm (Western Illinois) in tapered aluminum-alloy
posts (Fig. 1(a) and (b)). These are typical slender systems, with
reduced mass and low damping. As a result, approximately 140
units failed during this event.

Corresponding address: Department of Civil and Environmental Engineering, Northeastern University, 400 Snell Engineering Center, 360 Huntington
Avenue, Boston, MA 02115, USA. Tel.: +1 617 373 5186; fax: +1 617 373
4419.
E-mail address: lucac@coe.neu.edu (L. Caracoglia).

c 2006 Elsevier Ltd. All rights reserved.


0141-0296/$ - see front matter
doi:10.1016/j.engstruct.2006.06.023

The structures had been designed in accordance with code


recommendations. The occurrence of an unusual event or
configuration was therefore considered as the main suspect,
responsible for this incident. An extended investigation focused
on potential mechanisms involving wind-induced vibration
was conducted in order to consider the different aerodynamic
and aeroelastic phenomena capable of leading to multiple
system failures. The final objective of the research was not
only the identification of the most plausible causes but also
the evaluation of potential mitigation strategies to be directly
applied to the existing structures and that could be incorporated
in future applications.
Therefore, an experimental campaign was designed and
carried out to better address these issues, to identify the
mechanical characteristics responsible for the interaction with
the wind and to analyze potential solutions for the reduction of
the amplification.
2. Description of the problem
Aluminum-alloy tenon top poles (Fig. 1(a) and (b)) have
been introduced on major highway interchanges in Illinois. The

822

L. Caracoglia, N.P. Jones / Engineering Structures 29 (2007) 821831

Fig. 1. 13.7 m aluminum-alloy highway light pole. (a), (b) Example of collapsed units near Galesburg, Illinois, USA, 2003 (courtesy: Illinois Department of
Transportation). (c) Full-scale laboratory experiments.

main advantage consists in a lower mass (about three times


lighter than steel); however, higher flexibility results (E al =
70 GPa vs. E steel = 205 GPa). These are also usually lowdamping systems. The combination of lower mass with low
damping increases susceptibility to wind-induced vibration.
During a winter storm on February 11, 2003, approximately
140 tapered aluminum light poles in western Illinois collapsed
(at about the same time) at seven different interchanges on I74, US-150 and US-34 in Galesburg, IL [2]. Installation dates
of the poles in the Galesburg area were variable: most poles
were installed in 2000, in 1997 and others in 1999 [2]. At the
perceived moment of the failure the recorded wind speed was
about 22 m/s from the NW (steady winds with low turbulence).
Heavy snow was also present, with the temperature constantly
below zero degrees Celsius since the day before the collapse.
These large oscillations induced failure either in the breakaway supports (anchorages to ground, most common reported
cause) or, in many cases, a fracture failure initiated at the
handhole access in the pole itself. Poles were designed to
withstand high winds in accordance with US specifications [1],
and also equipped with canister vibration damper, located at
approximately two-thirds of the height L with respect to the
base, tuned to the second mode of vibration, and designed
to mitigate dynamic amplifications due to Karman vortex
shedding [35].
Three main sources of excitation were identified: alongwind buffeting (dominated by first mode), vortex-induced
vibration (at lock-in), and across-wind galloping [5], possibly
initiated by the presence of a buildup of frozen precipitation
on the surface of the pole or the luminaire. In addition, the
potential combination of one or more of the above phenomena
was considered. Both maximum capacity of the material,
and fatigue due to long-term cyclic loading (if applicable)
were investigated. For material capacity, the study focused

attention on the response of the break-aways since a direct


correspondence of load/stresses induced at the base of the
pole was readily possible; on the contrary, collapses involving
the pole cross-section were more difficult to characterize
and summarize due to an inherent variability in the failure
configuration (Fig. 1(a) and (b)).
3. Wind environment
Wind velocity is usually decomposed into a mean wind
component, associated with long-term variability in the flow
(10 min to 1 h) and a high-frequency component (order
of seconds, turbulence), usually responsible for the dynamic
structural response. The first component (10 min averaged wind
speed, U ) can be represented, for neutrally stratified stable
boundary-layer flows, as a function of the height with respect
to ground z, as [3]
U (z) = 2.5u ln (z/z 0 ) ,

(1)

with u the surface friction velocity and z 0 the roughness


length. Estimates of u and z 0 are derivable for different terrain,
site topography [6,7] and exposure category [1,8].
The turbulence-induced zero-mean wind fluctuations in the
along-wind, u, across-wind v and vertical w directions, are
responsible for the dynamic response. The Gust Velocity Factor
(GVF), G u [1], is defined as the ratio between the maximum
expected peak value of the velocity, U (z), and the averaged
value of the along-wind component U (z) through Eq. (1), i.e.,
G u (z) = U (z)/U (z).

(2)

The reference design value of wind speed (AASHTOIV, [1]) is defined as the three-second gust with fifty-year
return period at z = 10 m (basic wind speed, BWS). In the

L. Caracoglia, N.P. Jones / Engineering Structures 29 (2007) 821831

analyses the Characteristic Wind Speed (CWS), referred to


10 min averages, was employed. For comparison purposes of
CWS with BWS, the GVF was computed for exposure category
C open terrain with scattered obstructions, and estimated as
1.31 < G u < 1.41 [9]. For this area (exp. C) a BWS of 40 m/s
is prescribed [1], i.e., a CWS of 2830 m/s.
4. Strength failures associated with buffeting
Buffeting analyses were based upon the closed-form method
proposed in Ref. [10] for the study of masts and monotubular
steel towers due to fluctuating wind loading under stationary

wind conditions. The generic peak response effect R(z)


at a
specific location z, is referred to its corresponding mean value

R(z)
through an appropriate and unique Gust Effect Factor
(GEF) [11],

G R (z) = R(z)/
R,

(3)

depending on the location z and the quantity that is monitored


(for example, maximum along- or across-wind displacement,
moment, shear, stress at a particular cross section), and
including fluctuations in the wind energy content at different
frequencies (background response), dynamic amplification due
to the interaction with the structure (resonant response) and
partial loss of correlation of the pressures due to the span-wise
extension of the system. First-mode dominated response was
considered.
The dynamics characteristics of the poles were represented
by considering symmetry in the along-wind, x, and acrosswind, y, directions, i.e., a set of j independent mode shapes,
j,x (z) and j,y (z), with the same frequency but perpendicular
orientation. Although the asymmetry in the luminaire fixture
can produce twist and bending, torsional components were
neglected since the eccentricity was considered as secondary
in the analyses. Frequencies and modes were derived in closed
form by simulating the system as a cantilever beam with
tapering cross section and the representation of the structural
mode shapes in terms of Bessel functions [1214]. Tapered
light poles (13.7 m) with circular cross section and variable wall
thickness, D (between 5 and 7 mm), had been installed. The
outer diameter, linearly variable with height, was the same for
all cases: 152 mm at the top of the structure and 254 mm at the
base. Different luminarie models were present with mass M L
estimated between 18 kg and 27 kg.
Typical values of the first- and second-mode frequency for
D = 7 mm and luminaire mass M L equal to 0, 13 kg and
18 kg, respectively, are: f 1,x = 1.49 Hz, f 2,x = 7.61 Hz;
f 1,x = 1.24 Hz, f 2,x = 6.52 Hz; f 1,x = 1.17 Hz, f 2,x =
6.31 Hz. For thinner poles ( D = 5 mm) the frequencies
are: f 1,x = 1.49 Hz, f 2,x = 7.61 Hz; f 1,x = 1.18 Hz,
f 2,x = 6.35 Hz; f 1,x = 1.10 Hz, f 2,x = 6.14 Hz. The
quantity M L = 18 kg corresponds to the luminaire mass of
the models installed in the field, M L = 13 kg to that used in the
laboratory experiments (Section 9 and Fig. 1(c)). The first mode
frequency is located in the neighborhood of 1.0 Hz, tending to
0.9 Hz for larger M L . Analytically-derived frequency and mode

823

shapes were also compared with finite-element simulations; the


correspondence was very good [15].
Structural modal damping of the poles (percentage with
respect to critical) was assumed as 1,x = 0.5%, compatible
with similar metal units [16], as later confirmed by the
experimental results (Section 9). Normalized aerodynamic
force coefficients (drag, lift, C Dn , C Ln with n = 0 pole,
n = 1 luminaire) were specified in accordance with the
values indicated by AASHTO-IV. A constant value of C D0 =
1.1 independent of the Reynolds Number Re was considered
in this section, as a conservative value with respect to the
code, but compatible with the results for smooth cylinders at
Re = 250,000 [3]. This value is located in the proximity of
the range where the critical transition from laminar to turbulent
flow occurs, characterized by reduced values of drag; effects
will be investigated in the following sections. Aerodynamic
coefficients for the luminaries were estimated by comparison
with equivalent rectangular sections and without specific wind
tunnel tests.
The presence of both along-wind and across-wind buffeting
was simulated by independently computing the first mode
along-wind and across-wind response (and equivalent dynamic
loading) and combining the two effects by means of an
appropriate combination rule (non-simultaneous occurrence of
the peaks in the xz and yz planes, [10]). The Solari wind
spectrum was used [17] with meteorological and topography
conditions at the time of failure simulated through u =
1.4 m/s and z 0 = 0.02 m.
Two buffeting response characteristics were considered:
the maximum peak equivalent displacement due to wind
fluctuations in correspondence with the top of the pole (z = L,
with L height of the pole) and the maximum buffeting-induced
overturning moment and shear forces at the base, in particular
compared to the maximum tensile capacity of the anchorages
to ground (break-aways) determined from laboratory testing.
The failure criterion (wind condition necessary to exceed
the tensile capacity of anchorages) was selected in accordance
with a collapse limit state methodology and the most vulnerable
anchorage in both shear and tension. Dead load of the pole
with the potential presence of precipitation accumulation
was negligible. Two shear/bending moment mechanisms were
considered: a first mechanism in which all the four bolts
(Fig. 1(b)) are active and a second, more stringent mechanism,
in which two inactive bolts lie on the neutral axis.
An example of the simulation results is included in
Table 1(a) and (b), where the first-mode buffeting response for a
13.7 m pole with D = 5 mm and different wind characteristics
(u , z 0 ) is summarized. Similar studies are available for thicker
units. These tables show the maximum peak displacement at the
top of the pole X eq,Q (L) = X (L) + xeq,Q (L) (Q is labeled as
x, y or C equivalent combination if both xz and yz loads
are considered), total modal damping 1,Q , gust effect factors
G Q , root-mean-square and maximum dynamic displacements
in the two orthogonal planes xstd , xmax and ystd , ymax (with
Y = 0). All quantities are computed for z = L. In Table 1(a) the
effect associated with the variation of the surface roughness is
analyzed, while the failure criteria are investigated in Table 1(b)

824

L. Caracoglia, N.P. Jones / Engineering Structures 29 (2007) 821831

Table 1
First-mode gust buffeting response for a 13.7 m aluminum-alloy pole with D = 5 mm and 1,Struct = 0.5%. (a) u = 1.4 m/s and variable z 0 ; (b) U (10) = 22 m/s,
variable z 0 and u

1st (- -) and 2nd ( - ) failure mechanisms for variable geometric tolerance.

(moment, shear at the base indicated as M0C max , V0C max and
derived as equivalent static loads [10]).
For an average wind and turbulence scenario (z 0 = 0.02 m,
u = 1.4 m/s, U (10) = 22 m/s, along-wind turbulence
intensity Iu = 0.16) X eq,C (L) can vary in the range 0.45
0.60 m (Table 1(a)), corresponding to a dynamic oscillation
with 0.50 < xeq,C (L) < 0.60 m peak-to-peak amplitude
(Table 1(b)), and a mean deflection equal to 0.26 m under steady
wind for D = 5 mm. The first-mode total damping in the
along-wind direction, 1,x,tot = 1,x + 1,x,Aerod (structural
and aerodynamic components) is significant and mainly related
to the beneficial effect of the luminaire: 4%5% at U (10) =
22 m/s. A further parametric investigation, also considering
variability of C D0 and C D1 , suggests a minimum value of 3.4%.
Despite the moderate amplitudes, the moment and shear forces
recorded at the base are approximately two to three times lower
than the values necessary to induce failure in the break-aways.
In the tables, the wind conditions associated with
both failure mechanisms (dashed line for mechanism one,
dasheddotted line for mechanism two) were computed,
including the effect of the potential presence of geometric
tolerance in the installation of the anchorages to ground. It was
concluded that the wind characteristics or the extremely large
vibration necessary to exceed the tensile capacity of the anchor
bolts were not consistent with the meteorological conditions at
the time of the event, requiring much higher winds or physically
unreasonable turbulence intensity levels.
5. Influence of vortex shedding on the response during
collapse
Vortex shedding in the across-wind direction was neglected
since it usually involves higher modes, and dynamic interaction
can be excluded due to evident effectiveness of the canister
vibration damper (Section 2) tuned to the second mode. Lockin from vortex shedding was not responsible for failures,
evidenced by the observed large-amplitude oscillations and the

fact that the Strouhal relationship,


S = f Dref /U,

(4)

suggests that for S = 0.2, U (10) = 22 m/s and Dref =


203 mm (average value of outer diameter), f = 22 Hz
(i.e., third to fourth mode).
6. Fatigue failures derived from long-term exposure to
buffeting or lock-in conditions
Fatigue failures due to wind buffeting were also investigated
by means of the analytical method recently proposed in [18]
for the study of cylindrical vertical structures affected by wind
loading. The mechanical characteristics of the material were derived from AASTHO-IV, as integrated by recent studies [19,
20].
Wind directionality effects, usually beneficial in the
redistribution of stresses during the lifetime of the structure,
were neglected. Winds with constant direction were considered
with 20% probability of absence of wind [18]. A full
description of the procedure can be found in [15]. Simulations
indicated that the fatigue life of the breakaways could be
conservatively estimated as about eight years, depending on
the failure mechanism and considering the effective action of
a limited number of anchorages (two) due to wind direction.
Fatigue life for the aluminum cross-section at the base was also
estimated as about seven to eight years, by taking into account
different reduction factors in the definition of maximum fatigue
capacity for aluminum. The influence of the drag coefficient in
the critical range of Re was also carefully analyzed.
In general, these predictions are extremely sensitive to the
input parameters, such as the statistical distribution of the wind
speed and the reference fatigue curve of the material (maximum
differential stress vs. load cycles). In particular, estimates can
differ by one order of magnitude due to the nonlinear nature of
the fatigue curves; this fact was observed through a parametric
investigation conducted by varying the material reduction factor

L. Caracoglia, N.P. Jones / Engineering Structures 29 (2007) 821831

(fatigue limit states) either applied to the aluminum crosssection (pole base) or to the break-aways tensile strength
to simulate the effects associated with the different fatigue
categories (severity of the loading). In a recent study [21] it was
shown that fatigue life of steel poles can be as low as one year
under special wind exposure conditions and due to the effects
of drag reduction in the critical Re region. The subsequent
application of the closed-form method proposed in [21] to
this specific case suggested a life expectation variable between
two and seven years for the aluminum pole base cross-section,
although a penalizing factor relating stress variation to wind
speed was employed due to the limited information and data.
Fatigue-life predictions larger than seven to eight years
seem compatible with previous results derived for steel poles
through the method utilized in [18]. Moreover, buffetingrelated fatigue did not seem compatible with the multiple
simultaneous occurrences experienced in this particular event,
also considering the fact that most units were installed between
1998 and 2000. Specifically, a concurrent collapse of systems
located in the same area with different failure characteristics
(and service ages) exclusively associated with fatigue was
considered as statistically improbable. Fatigue issues related to
vortex shedding were also excluded due to the extremely low
value of lock-in velocity range corresponding to the first mode,
the presence of the impact canister damper and the indication
of large-amplitude vibration at the time of the incident.
7. Across-wind galloping instability
Galloping is an aeroelastic instability phenomenon involving
large-amplitude oscillation in the across-wind direction [3,
5], triggered by a decrement of the aerodynamic damping
component and involving energy transfer from the wind flow.
The necessary condition for across-wind first-mode galloping
is
1,y,tot = 1,y + 1,y,Aerod = 0.

(5)

The vanishing of the total damping in Eq. (5) can be interpreted


as the boundary between stable (1,y,tot > 0) and unstable
dynamic oscillation (1,y,tot < 0). Linearized analysis and a
quasi-steady approach can be adopted since typical frequencies
of the poles were much lower than those associated with vortex
shedding for U = 22 m/s [5]. Since 1,Aerod is a function
of the wind velocity, Eq. (5) can be translated into a critical
wind velocity for galloping onset for a single degree of freedom
system [3,5], as
UCR / ( f 1 Dref )
= 4

C L0

+ CD

1



1 
2
2 m 1,eq 1,y Dref
,

(6)

where f 1,y is the across-wind first-mode frequency of the


system, Dref is a reference dimension of the cross section (at
z = 0.6L), m 1,eq is an equivalent first-mode mass per unit
length (including the presence of the luminaire), is the air
density, 1,y is the structural modal damping. The mean wind
velocity UCR was referenced to z = 10 m since the presence of

825

a boundary-layer profile was considered; turbulence fluctuation



effects on Eq. (6) were neglected. The quantity C L0 + C D , a
combination of the drag coefficient per unit length C D and the
first derivative of the lift coefficient with respect to the angle of
attack C L0 = C L /, must be negative (GlauertDen Hartog
Criterion).
Although symmetric circular sections are immune from
this type of phenomenon, the meteorological situation
corresponding to the investigated event suggests that the
potential presence of frozen precipitation on the surface of
the pole could not be excluded. In particular, precipitation
such as rain or snow close to freezing temperature, can supercool at low temperatures when falling from the upper levels
of the atmosphere due to barometric pressure gradient. Super
cool water (liquid below freezing temperature) freezes almost
immediately when reaching another surface characterized by a
significant temperature difference with respect to the air. This
suggests that slurry/freezing precipitation could have possibly
adhered to the surface of the pole.
Eq. (6) was cast into a more general formulation for
the analysis of first-mode galloping oscillations of poles and
vertical masts [10,22], as indicated in [23], by incorporating
modal quantities and simulating the presence of frozen
precipitation on the whole surface of the pole and in
correspondence with the top of the luminaire, separately.
Analyses included variable geometry and wind characteristics,
with 1,y = 0.5%.
Lift coefficients, C L0 and C L0 were derived in accordance
with the experimental data in [22] for a circular cross section
with ice thickness variable from 3% to 5% with respect to
the reference diameter of the body and a coverage area of approximately 120 . In addition, experimental studies [24] report
that some luminaire shapes are potentially prone to gallopingtype oscillation (in the absence of frozen precipitation). Attention was devoted to the study of the drag coefficient, C D0 , in
the critical range of Re. As an example, it was found that for
Re = 250,000 and for cylinders with dimensionless roughness
coefficient of the order of 104 to 103 (metal surface), C D0
can vary between 0.55 and 0.73 according to [25], and 0.85 [3],
the latter value also being suggested by the AASHTO-IV interpolating equation for a wind speed compatible with the meteorological conditions of the event. Since the analysis of a collection of different experimental tests [26] indicated considerable
variability of C D0 , a sensitivity investigation was conducted by
evaluating distinct limiting conditions with respect to a reference value C D0 = 1.0. Results are summarized in Table 2, for
a 13.7 m pole with M L = 18 kg, variable wind profile with
z 0 = 0.02 m and assuming modal damping and mass distribution in accordance with mode shapes 1,y .
For the basic value of C D0 = 1.0 the galloping critical
velocity UCR at z = 10 becomes extremely large when the
surface of the pole is covered by 3% ice, for an equivalent 120
coverage area from the top to the base (case (i) in Table 2). In
contrast, the critical wind speed derived for a 5% ice thickness
(1518 m/s cases (ii) and (iii)) is low for the pole that was
considered relative to the value at the time of collapse (22 m/s),
although the simulated deposit of frozen precipitation seems

826

L. Caracoglia, N.P. Jones / Engineering Structures 29 (2007) 821831

Table 2
First-mode galloping critical velocity, UCR , at z = 10 m, for a 13.7 m pole with M L = 18 kg as a function of wall thickness, D , and variable C D , C L , C L0

(i)
(ii)
(iii)
(iv)
(v)
(vi)
(vii)
(viii)

Case

D
(mm)

C D0

0
C L0

C D1

0
C L1

UCR
(m/s)

3% simulated ice on pole


5% simulated ice on pole
5% simulated ice on pole
Sim. ice on lum. (conservative)
Ice on lum. (conserv.), low C D0
Simulated ice on lum., low C D0
Simulated ice on lum., low C D0 , C D1
Simulated ice on lum., no C D0 , C D1

7
7
5
7
7
7
7
7

1.0
1.0
1.0
1.0
0.4
0.4
0.4
0.0

1.5
1.9
1.9
0.0
0.0
0.0
0.0
0.0

1.4
1.4
1.4
1.4
1.4
1.4
0.5
0.0

0.0
0.0
0.0
5.3
5.3
3.0
2.0
2.0

69
18
15
110
14
208
300
17

extremely penalizing and incompatible with the meteorological


conditions during the investigated event. This observation also
excludes the direct influence of the drag reduction and Re,
which would lead to extremely low velocity values (case (v)).
Moreover, galloping due to frozen precipitation exclusively
concentrated on the luminaire fixture was not a plausible
explanation, since the critical speed derived from the
simulations was either extremely high (case (iv) in Table 2) or
physically unreasonable (cases (vi) and (vii)). The simultaneous
or alternate vanishing of C D0 and C D1 led to estimates closer to
the target value, but this circumstance does not seem physically
possible (high surface roughness of the poles, and luminaries
are bluff bodies with large flow separation). However, partial
coverage of freezing rain, distributed on a selected portion
of the pole, perhaps combined with unstable luminaire shape
and affected by meteorological conditions, seemed consistent
with the wind and environmental conditions, i.e., similar to
cases (ii) and (iii) in Table 2. In any case, further experimental
and wind tunnel studies would be desirable to better ascertain
the aerodynamic coefficients of the complex luminaire crosssection and to exclude or confirm its influence on the event.
From Table 2 and the exclusion of other factors (Sections 4
6), it was concluded that the most plausible multiple-failure
cause could be related to across-wind galloping instability or,
possibly, an aeroelastic phenomenon initiated by the build-up of
freezing rain and characterized by large values of the resonant
portion of the dynamic response power spectral density [10].
This condition, equivalent to buffeting prior to galloping, is
schematically depicted in Fig. 2 and indicated by U R and
0,R /Dref (normalized root-mean-square response), whereas
the galloping onset corresponds to the asymptotic behavior at
UCR .
A subsequent analysis [23] demonstrated that this class of
aluminum poles (thin wall, L = 13.7 m, small luminaire
fixture) were the most vulnerable to this type of aeroelastically
enhanced phenomenon. Furthermore, this problem was typical
of aluminum units since it was also shown that for equivalent
steel poles the same behavior was associated with much higher
and physically unreasonable wind velocities.
8. Vibration mitigation
The recommendations for vibration reduction were proposed
by recalling the definition of the instability criterion for windinduced vibration of cylinders and masts, according to which

Fig. 2. Schematic representation of galloping threshold UCR and aeroelastically enhanced phenomenon as a function of normalized root-mean-square
across-wind response.

the equivalent first-mode dimensionless Scruton Number,



1
2
K S = m 1,eq 1,y Dref
,

(7)

must satisfy the condition K S > N (see Eq. (6)). This quantity
is a measure of susceptibility to aerodynamic instability
(typically, lock-in from vortex shedding). The positive quantity
N varies from case to case; however, for masts with roughness
or longitudinal attachments in the critical Reynolds Number
regime, a value of 4.0 is suggested in [27], while a K S between
1.2 and 1.4 was computed for the analyzed case, for a reference
diameter (0.2 m), equivalent mass corresponding to first-mode
vibration and 0.5% damping.1
An increment of structural damping was strongly indicated
as a possible solution, leading to an increase of galloping
instability threshold. Different damping devices have been
proposed for light poles, traffic signals and luminaries
susceptible to wind excitation [20,28], including, for example,
aerodynamic damping devices [29], TMD/impact dampers [30]
on long horizontal arms of traffic lights or impact ball dampers
on posts to reduce Karman-vortex-induced oscillation [31]. In
1 A recent development of this research, based upon a statistical method
for the numerical simulation of the failure probability ( p f ) due to galloping
induced by simulated precipitation deposit, suggested that the original threshold
indicated in this paper may be replaced by N = 3.0, approximately equivalent
to 104 < p f < 103 (yearly value) [23].

L. Caracoglia, N.P. Jones / Engineering Structures 29 (2007) 821831

this study a simple yet effective and inexpensive method for


the reduction of wind-induced vibration on vertical structures
was considered: a chain damper, commonly used on tall masts
or slender truss-type towers and sometimes on shorter vertical
systems.
The design of such devices was performed by employing
a series of optimal damping curves [32,33] derived for a
device suspended inside a container attached to a vibrating
structure, subject to forced vibration at angular frequency
and constant amplitude x0 . Vibration reduction is given in terms
of a normalized damping group, ceq (lm)1 ,
as a function of
the frequency ratio /Ch , with Ch = 1.22 g/l frequency
of the pendulum, chain length l, mass per unit length m, and
an equivalent viscous damping coefficient, ceq . For /Ch >
3 but below a critical value the damping group becomes
approximately constant (optimal) and independent of the
excitation frequency. This condition can be estimated as [32],

ceq /lm optimal = 4 2 d/x0 .
(8)
Eq. (8) also shows that the damping is amplitude dependent.
Different operational regimes can be observed, in which partial,
full or no impact occurs. The best performance is achieved
when two impacts per cycle occur. However, chain dampers
are not effective at very low amplitudes or, equivalently,
high frequency [32,34]. At large oscillation amplitudes the
number of impacts usually increases and becomes irregular,
while beyond a critical value of frequency ratio (depending on
d/x0 ), damping drastically diminishes (no impacts). Noise is
a disadvantage of the approach, although it can be mitigated
through a rubber covered chain without significant reduction in
the impact restitution coefficient [32].
An appropriately designed device (l = 3.05 m, m =
1.48 kg/m and /Ch
= 3) was employed, with estimated
modal damping equal to 7.0% for d/x0 = 1.0 (Eq. (8))
and 3.5% for d/x0 = 0.5, i.e., a first-mode Scruton Number
K S > 7.0 (depending on amplitude). The quantity x0 can
be interpreted as the amplitude associated with the firstmode generalized coordinate, i.e., an equivalent dynamic
displacement at the top of the pole.
9. Experimental investigation of the proposed damper
performance
Due to the inherent uncertainty associated with the definition
of the impact damper performance an experimental campaign
was conducted in the NSEL (Newmark Structural Engineering
Laboratory, University of Illinois at Urbana-Champaign, USA)
to study the dynamic response characteristics of aluminum
tapered poles and validate the proposed solution. Prior to the
installation of the device, free-vibration tests were performed in
order to extract the frequency and the damping characteristics
of the investigated pole, a 13.7 m tapered straight pole with
D = 5 mm, similar to those involved in the multiple collapses.
The pole was originally equipped with canister vibration
damper (Section 2) tuned to the second mode. Fig. 1(c) shows
some of the installation and experimental stages; a luminaire
with M L = 13 kg was used.

827

A vertical orientation was preferred due to the versatility


of the apparatus and the necessity of testing the proposed
mitigation device. The specimen was anchored to a heavy
concrete block, located in the basement of the laboratory,
through an opening in the strong floor. The testing involved
different configurations and combinations: with and without the
luminaire, in both the absence and the presence of the chain
impact damper. Also, the interaction between canister damper
and the new device was considered.
Instrumentation included in-plane (AC1, xz plane of the
luminaire in Fig. 1(c)) and out-of-plane (AC2, yz plane and
orthogonal to xz) single-channel accelerometers at the top
of the unit, one in-plane accelerometer (AC3) located at z =
0.56 L = 7.6 m, close to the anti-node of the second mode
and a displacement transducer at z = 11.8 m in the x direction
(Fig. 1(c)), referenced to an existing reaction wall.
During the free-vibration tests the structure was quasistatically pulled backward in the x (in-plane) direction through
a hydraulic actuator, anchored to the floor of the laboratory
(z = 4.46 m with respect to the pole base) and released
from this initial condition. The corresponding initial deflection
at the top of the pole, x0,L , was derived from the available
instrumentation. The range of initial conditions 0.06 < x0,L <
0.30 m was considered to study the amplitude-dependent
response of the chain.
Experimental studies conducted on steel mono-tubular
masts, directly in the field [16], suggest that structural damping
of such systems is variable and can be a function of several
parameters including, for example, the component related to
soil-structure interaction at the base/foundation of the pole.
Climate and temperature (such as, for example, the reported
conditions at the moment of collapse) may have detrimental
effects on such values. Unfortunately, direct experimental
simulation of these effects was not possible.
The laboratory tests confirm that aluminum poles are
structures where the inherent damping component is low and
that the introduction of a damper was desirable to reduce the
sensitivity to aeroelastic excitation. As an example, Fig. 3(a)
and (b) show the time histories associated with the freevibration test no. 01BN040, corresponding to the case without
luminaire, after the removal of the canister damper and x0,L =
0.13 m (the minus sign denotes deflection due to backward
pulling), i.e., for moderate oscillation amplitudes. Original data
(in-plane displacement at z = 11.8 m) were digitally bandpass filtered to represent the response associated with the
fundamental mode without (a) and with (b) chain damper.
In the absence of chain damper (Fig. 3(a)) an initial
oscillation decay is followed by an increment at t = 25 s
and subsequent second decrement. This phenomenon is not
associated with the inherent damping of the pole but with the
presence of a false beating effect, observable even in the
absence of the eccentric luminaire. An imperfect planarity of
the x and y modes, j,x and j,y , most probably related to nonhomogeneity and asymmetries of the pole in the xz and yz
planes, is responsible for a two-dimensional response. Closelyspaced but distinct frequencies (1%2% difference) were also

828

L. Caracoglia, N.P. Jones / Engineering Structures 29 (2007) 821831

Fig. 3. Example of 13.7 m pole free-vibration tests (M L = 0, canister damper


removed); time history recorded at z = 11.8 m (x0,L = 0.13 m before
release); (a) bare pole, (b) with chain damper (band-pass filtered signal).

Fig. 4. In-plane (AC1) and out-of-plane (AC2) first-mode filtered accelerations


(13.7 m pole, M L = 0 and canister damper removed), x0,L = 0.13 m before
release; (a) bare pole, (b) with chain damper.

recorded. No significant first-mode response differences were


observed before and after removing the canister damper.
Fig. 3(b) is related to the same configuration as in Fig. 3(a)
with the addition of the chain damper. Amplitude rapidly
decreases after a few cycles and the two-dimensional response
of the pole is reduced, due to the effectiveness of the passive
control device. Regular impacts (two per cycle) were noticed
in most cases for a relatively wide range of amplitudes, also
confirming the predictions (Section 8). Nevertheless, during the
free-decay evolution the pole typically undergoes elliptical or
circular trajectories with moderate oscillation amplitudes for a
limited duration, during which the effectiveness of the chain is
drastically reduced (whirling motion).
An example of a two-dimensional trajectory in the absence
and presence of chain damper is presented in Fig. 4(a) and (b),
respectively corresponding to Fig. 3(a) and (b). Acceleration
time-histories associated with the fundamental mode (after
band-pass filtering) are depicted, and restricted to the interval
5 < t < 20 s. In Fig. 3(b) the two-dimensional motion is
suppressed due to the high levels of damping provided by the
hanging chain.
The coupling of the chain with the existing passive device
was also analyzed; one of the perceived advantages is related to
the promotion of energy transfer from the fundamental to higher

modes, which can be dissipated through the canister damper.


The chain damper alone is not effective when the response
is dominated by second or higher modes, in consideration of
its location and the presence of amplitude-dependent critical
frequency ratios [32]; therefore, the combination of both
devices is desirable.
Accurate damping estimates in the absence of the chain were
particularly difficult due to the presence of large-amplitude twodimensional motion (Fig. 4(a)) and the interaction between x
and y response components. The separation of the two modes
by signal filtering was not possible. Alternative time-frequency
multi-scale techniques (e.g., [35,36]) were employed but the
extraction of damping ratios through mode decomposition
was still challenging and influenced by the closeness of the
modes. In these cases damping was evaluated from the residual
oscillation in the out-of-plane direction extracted from AC2 (ydirection), usually observed during the tests with chain damper,
since the latter is ineffective at low amplitudes. A typical value
derived for M L = 0 and for the out-of-plane mode ( f 1,y =
1.418 Hz, measured) is 1,y = 0.2%.
Fig. 5 presents an example of a first-mode damping
estimate for M L = 0 ( f 1,x = 1.417 Hz, measured), with
x0,L = 0.24 m before release and in the presence of
chain damper, from in-plane displacement time history at

829

L. Caracoglia, N.P. Jones / Engineering Structures 29 (2007) 821831

Table 3
First-mode structural damping (1,x ) of a 13.7 m aluminum alloy highway pole
for large (a) and moderate (b) oscillation amplitudes
(a)
Case

f (Hz)
(note 2)

x0,L
(m)

1
(%)

M L = 0, NO chain
(test01BNC081, note 1)
M L = 0, chain
(test01BNC081)
M L = 13 kg, NO chain
(test01BC080, note 1)
M L = 13 kg, chain
(test01BC080)

1.418
(1.461)
1.417

n.a.

0.2

0.24

4.1

1.157
(1.156)
1.109

n.a.

0.1

0.25

2.1

1.418
(1.468)
1.412

n.a.

0.1

0.12

3.9

1.158
(1.158)
1.142

n.a.

0.1

0.12

2.4

(b)
M L = 0, NO chain
(test01BNC040, note 1)
M L = 0, chain
(test01BNC040)
M L = 29 Lb, NO chain
(test01BC040, note 1)
M L = 13 kg, chain
(test01BC040)

Note 1: 1,y derived from residual values in the presence of hanging chain.
Note 2: Measured f in parentheses computed during free-vibr. without chain.

Fig. 5. Example of damping assessment from free-vibration tests (13.7 m pole


with chain damper; M L = 0, canister damper removed, x0,L = 0.24 m
before release). (a) In-plane displacement time-history at z = 11.8 m, (b)
damping ratio estimates (least squares) through envelope curve, A(t).

z = 11.8 m (digitally band-pass filtered). From least-squares


linear regression in the semi-logarithmic plane of A(t) (timevarying amplitude derived through Hilbert transform [37]) an
equivalent viscous damping corresponding to the selected mode
can be derived. This method is sensitive to the time interval
(between 8 and 20 s in Fig. 5(b)) selected for the identification
of A(t). The equivalent estimated curve corresponding to linear
damping is also indicated in Fig. 5(b) (dashed thick line);
deviations were linked to the non-linearity of the device. This
example corresponds to large oscillation amplitudes (0.38 m
peak to peak at z = 11.8). From the analysis of the experimental
envelope curve (solid line), a kink is evident at t = 13 s,
for amplitudes A = 0.03 m at z = 11.8 m, associated with
the limit beyond which no impacts are recorded. Although
the procedure allowed for the derivation of damping as a
function of amplitude, it was decided to conservatively evaluate
the performance of the chain from a design perspective by
considering an averaged value accounting for both regions
of active (t < 13 s in Fig. 5(b)) and inactive chain (t >
13 s). In the figure, the computation suggests 1,x = 4.1%,
approximately ten times larger than the original value. Other
techniques for the derivation of 1 were analyzed such as
a method based upon a direct nonlinear regression of either

acceleration or displacement records, substantially confirming


the original estimates.
Table 3(a) and (b) summarize the results related to the
computation of first-mode damping through Hilbert transform
for different pole configurations without canister damper (tests
conducted before the removal of the canister damper confirmed
the findings reported in these tables). Equivalent viscous
damping was derived for the classes of large (Table 3(a)) and
moderate (Table 3(b)) oscillation amplitudes. Time intervals for
the calculation of 1 were selected in order to be representative
of large (x0,L > 0.2 m) and moderate (x0,L < 0.1 m)
vibration amplitudes. In the table, the distinction between x
and y components is omitted. Values of 1,y without first-mode
damper were derived from residual behavior in the presence
of hanging chain, as described above. In these cases amplitude
dependence is not applicable (n. a.). Results confirm low values
of first-mode structural damping component, between 0.1% and
0.2%. Differences can be attributed to the fact that residual
oscillation regimes (inactive chain) are influenced by the initial
conditions and the decay trajectories linked to the coupling in
the x and y directions.
A cross comparison of the two tables and for equivalent
x0,L suggests limited amplitude effects. An overall reduction
of damping is observed in the presence of the luminaire at
both moderate and large amplitudes with respect to the case
with M L = 0. This effect was interpreted by the fact that
the eccentric mass at the top of the pole promotes elliptic
trajectories and, as a consequence, the whirling of the chain
may be more quickly initiated. In general, experiments suggest
an equivalent 1,x located between 2.5% and 4.0% depending
on the different configurations and amplitudes. It is worth
recalling that such values were conservatively derived as shown

830

L. Caracoglia, N.P. Jones / Engineering Structures 29 (2007) 821831

in Fig. 5(b), in which the inactive-chain behavior is accounted


for in the calculation of the equivalent damping ratio.
10. Concluding remarks
This paper presents the results of a recent analytical,
numerical and experimental investigation focused on the
analysis of multiple collapses experienced by highway light
poles during a wind storm in Western Illinois, USA. The
study was performed in order to identify the potential causes
associated with the failures. It was concluded that, although
the poles were designed according to standard specifications,
an unusual event, in which the combination of wind and
frozen precipitation was observed, could be responsible for
large vibration amplitudes (aeroelastically related or enhanced).
This conclusion was derived after carefully analyzing other
possible causes of failure, including collapse limit states
governed by along-wind buffeting, fatigue cracking associated
with long-term exposure to turbulence buffeting in the critical
Reynolds range (more often indicated as a primary concern
for these systems) and von-Karman vortex shedding. All these
mechanisms were excluded since they did not seem compatible
with the particular meteorological condition and, at the same
time, with the failure modes recorded in the field.
Since the nature of the oscillations was clearly affected by
insufficient structural damping in the fundamental mode, the
proposed mitigation solution considered the installation of an
impact damper (hanging chain), optimized for the first mode,
as an inexpensive yet effective method for drastically reducing
the susceptibility to wind excitation.
The performance of the damper was evaluated through an
experimental campaign that confirmed a significant increment
in the damping ratio. The study showed that testing is crucial
in these situations, despite the relatively simple structure
due to the inherent variability in the configuration, evident
during the laboratory tests, and the uncertainty related to
the exact characterization of field situation (environmental
loads, geometric and physical characteristics of the units).
In addition, despite the limits on structural damping at large
amplitudes and the presence of three-dimensional motion
effects partially neutralizing the effects of the proposed damper,
the compatibility of the free-vibration decay curves with the
desirable or required damping levels persists. Moreover, the
investigation demonstrates that, in the case of impact dampers
the assessment of the performance needs to carefully address
the amplitude-dependent behavior.
Finally, this study and supplemental investigations concluded that, in general, aluminum-alloy light poles, although
vulnerable to wind-induced failures associated with this particular event, can be employed as an alternative to steel units, if appropriately designed and without the addition of the proposed
first-mode damper depending on the geometry and configuration (e.g., thicker and shorter units) [23].
Acknowledgements
This research was supported by the Illinois Department
of Transportation (Grant IHR-R27, 20032004) under the

technical direction of Mr. David L. Lippert of the Bureau of


Materials and Physical Research. Experiments were performed
at the University of Illinois, Urbana-Champaign, Newmark
Structural Engineering Laboratory (NSEL). The collaboration
of Dr. Grzegorz Banas and Mr. Timothy J. Prunkard, technical
supervisors in the NSEL and their staff was much appreciated
during the design and development of the tests.
Any findings do not necessarily reflect the official view,
opinions and recommendations of the Illinois Department of
Transportation, the United States Department of Transportation, or Federal Highway Administration.
References
[1] AASHTO. Standard specifications for structural supports for highway
signs, luminaires and traffic signals. 4th ed. Washington (DC, USA):
American Association of State Highway and Transportation Officials;
2001.
[2] IL-DOT . Background information regarding the luminaire failures on the
I-80 Le Claire Bridge over the Mississippi River and the pole failures in
Galesburg and Woodhull Areas. Documentation transmitted to UIUC on
August 21, 2003. Illinois Department of Transportation; 2003.
[3] Simiu E, Scanlan RH. Wind effects on structures. New York (NY, USA):
John Wiley and Sons; 1996.
[4] Goswami I, Scanlan RH, Jones NP. Vortex shedding from circular
cylinders: Experimental data and a new model. Journal of Wind
Engineering and Industrial Aerodynamics 1992;41(13):76374.
[5] Blevins RD. Flow-induced vibration. 2nd ed. New York (NY, USA): Van
Nostrand Reinhold; 1990.
[6] Wieringa J. Updating the Davenport roughness classification. Journal of
Wind Engineering and Industrial Aerodynamics 1992;41(13):35768.
[7] Tieleman HW. Roughness estimation for wind-load simulation experiments. Journal of Wind Engineering and Industrial Aerodynamics 2003;
91(9):116373.
[8] ASCE. Minimum design loads for buildings and other structures (ASCE702). Reston (VA, USA): American Society of Civil Engineers; 2002.
[9] Solari G, Kareem A. On the formulation of ASCE7-95 gust effect factor.
Journal of Wind Engineering and Industrial Aerodynamics 1998;7778:
67384.
[10] Pagnini LC, Solari G. Gust buffeting and aeroelastic behaviour of poles
and monotubular towers. Journal of Fluids and Structures 1999;13(78):
877905.
[11] Piccardo G, Solari G. 3-D gust effect factor for slender vertical structures.
Probabilistic Engineering Mechanics 2002;17(2):14355.
[12] Mabie HM, Rogers CB. Transverse vibrations of double-tapered
cantilever beams. Journal of the Acoustical Society of America 1972;
51(2):17714.
[13] Mabie HM, Rogers CB. Transverse vibrations of double-tapered
cantilever beams with end support and with end mass. Journal of the
Acoustical Society of America 1974;55(5):9869.
[14] Gorman DJ. Free-vibration analysis of beams and shafts. New York (NY,
USA): John Wiley and Sons; 1975.
[15] Caracoglia L, Jones NP. Analysis of light pole failures in Illinois.
Research Report. Urbana (IL, USA): University of Illinois at UrbanaChampaign; 2004.
[16] Pagnini LC, Lagomarsino S, Solari G. Experimental assessment of the
damping of steel poles and monotubular towers. Costruzioni Metalliche.
Journal of the Italian Technical Association for Steel Construction (CTA)
1999;(LI(1)):3951 [in Italian].
[17] Solari G. Turbulence modeling for gust loading. Journal of Structural
Engineering 1987;107(7):155069.
[18] Repetto MP, Solari G. Dynamic alongwind fatigue of slender vertical
structures. Engineering Structures 2001;23(12):162233.
[19] Kaczinski MR, Dexter RJ, Van Dien JP. Fatigue-resistant design of
cantilevered signal, sign and light supports. NCHRP Report 412.
Washington (DC, USA): Transportation Research Board - National
Research Council; 1998.

L. Caracoglia, N.P. Jones / Engineering Structures 29 (2007) 821831


[20] Dexter RJ, Ricker MJ. Fatigue-resistant design of cantilevered signal,
sign and light supports. NCHRP Report 469. Washington (DC, USA):
Transportation Research Board - National Research Council; 2002.
[21] Robertson AP, Holmes JD, Smith BW. Verification of closed-form
solutions of fatigue life under along-wind loading. Engineering Structures
2004;26(10):13817.
[22] Durgin FH, Palmer DA, White RW. The galloping instability of ice coated
poles. Journal of Wind Engineering and Industrial Aerodynamics 1992;
41(13):67586.
[23] Caracoglia L, Jones NP. Wind-induced failures of highway light poles
during winter storms. In: 6th ASME international symposium on fluidstructure interaction, aeroelasticity, flow-induced vibration and noise,
2006 pressure vessels and piping division conference. Vancouver (BC,
Canada): American Society of Mechanical Engineers; 2006. Paper
PVP2006-ICPVT11-93579.
[24] Barnard RH. Wind loading and instabilities of catenary-suspended road
lamps. In: Cook NJ, editor. Wind engineering: First IAWE European
and African regional conference. London (UK): Thomas Telford; 1993.
p. 4518.
[25] EC-3. Design of steel structures, eurocode 3, part 1-1: General rules and
rules for buildings. European Committee for Standardisation, 1993.
[26] Barre C, Barnaud G. High Reynolds number simulation techniques and
their application to shaped structures model test. In: Cook NJ, editor. Wind
Engineering: First IAWE European and African regional conference.
London (UK): Thomas Telford; 1993. p. 8393.
[27] Melbourne WH. Predicting the cross-wind response of masts and
structural members. Journal of Wind Engineering and Industrial
Aerodynamics 1997;6971:91103.

831

[28] Fouad FH et al. Structural supports of highway signs, luminaires


and traffic signals. NCHRP Report 494. Washington (DC, USA):
Transportation Research Board - National Research Council; 2003.
[29] Pulipaka N, Sarkar PP, McDonald JR. On galloping vibration of
traffic signal structures. Journal of Wind Engineering and Industrial
Aerodynamics 1998;7778:32736.
[30] Cook RA, Bloomquist D, Richard DS, Kalajian MA. Damping of
cantilevered traffic signal structures. Journal of Structural Engineering
2001;127(12):147683.
[31] Jo I, Kaneko T, Nagatsu S, Takahashi C, Kimura M. Development of
highway light pole with resistance to wind vortex-induced oscillations.
Technical Report 21. Kawasaki Steel Corporation; 1989.
[32] Reed WHI. Hanging-chain impact dampers: a simple method for damping
tall flexible structures. In: International research seminar wind effects
on buildings and structures. 1967, p. 283321.
[33] Koss LL, Melbourne WH. Chain dampers for control of wind-induced
vibration of tower and mast structures. Engineering Structures 1995;
17(9):6225.
[34] Masri SF. General motion of impact dampers. Journal of the Acoustical
Society of America 1967;47(1):22937.
[35] Staszewski WJ. Identification of damping in mdof systems using timescale decomposition. Journal of Sound and Vibration 1997;203(2):
283305.
[36] Kijewski T, Kareem A. Wavelet transforms for system identification in
civil engineering. Computer-aided Civil and Infrastructure Engineering
2003;18:33955.
[37] Bendat JS, Piersol AG. Random data analysis and measurement
procedures. 3rd ed. New York (NY, USA): John Wiley and Sons; 2000.

You might also like