You are on page 1of 19

Journal of the Chinese Institute of Engineers, Vol. 33, No. 4, pp.

597-615 (2010)

597

WIND-RESISTANT DESIGN OF HIGH MAST STRUCTURES

Ching-Wen Chien*, Jing-Jong Jang, and Yi-Chao Li

ABSTRACT
Today, a large number of High Mast Structures (HMS) is being constructed
around the world. Usually, the structural systems are presented as monotubular towers or masts with a mass at the tip which has utility equipment mounted. The
monotubular configuration of an HMS has a large ratio of height to horizontal
dimension. This means that the HMS is not only very slender, but also more windsensitive than any other common structures. Since failures of masts and monotubular
towers or poles have occurred often, this confirms the necessity for a better understanding of wind-excited behaviour on the HMS, and also a better design for a windresistant HMS. This paper illustrates the basic theories of behaviors of along-wind
response, and across-wind response. Prior to the present, design codes for HMSes
have not been standardized. Therefore, this study has aimed to develop a suitable
criterion of Wind Resistant Design (WRD) for HMS. We wish to produce guidelines
to be used when designing a wind-resistant HMS to counter wind-excited responses.
This paper starts with a theoretical approach. Computational Fluid Dynamics (CFD)
and Finite Element Method (FEM) are reviewed and applied to the analytical development of the design criteria. Then, the design criteria and the procedure of WRD are
established. And finally, a case study is presented for illustration. There are three
major findings in this study. First, the results of the analysis of CFD show that when
the polygon sides are more than 16, the total drag coefficient tends to the constant,
which closes to leeward drag coefficient, and the windward coefficient tends to be
zero. Second, its recommended that an HMS had better adopt a gust effect factor, G,
equal to 2.33 due to flexibility by geometric configuration and structural properties.
Third, the total maximum response limitation of an HMS should be considered in
across-wind analysis for WRD procedures.
Key Words: High Mast Structure (HMS), Wind-Resistant Design (WRD), Computational
Fluid Dynamics (CFD).

I. INTRODUCTION
The modern High Mast Structure (HMS) originated nearly at the beginning of the 20th century.
Typically, there is a functional utility on the tip of
the structure such as a wind-force turbine, a radio
wave transmitter, a radar, or some lamps and lanterns.
Up to today, different configurations of tubular
HMSes have been used widely in modern civil
*Corresponding author. (Tel: 886-2-26625858 ext. 58238; Fax:
886-2-26646102; Email: d93520010@mail.ntou.edu.tw)
C. W. Chien and J. J. Jang are with the Departemt of Harbor and
River Engineering, National Taiwan Ocean University, 2 Pei-Ning
Rd., Keelung 202, Taiwan, R.O.C.
Y. C. Li is with the Architecture and Building Research Institute,
Ministry of the Interior, Taipei, Taiwan, R.O.C.

construction. The unit costs of these structures are


low, however, their designs may cause problems
(Solaria, 1999). In the past years, many accidents
and much damage have been caused by strong wind
or wind-induced vibrations in such structures. These
frequent accidents confirm the necessity for a better
understanding about HMSs geometric configurations
and structural properties, and also the necessity to develop wind-resistant design criteria for HMS for the
use of designers.
We first focus on geometric configurations and
shape factors to compare the different codes. Then,
the structural and geometric properties of existing High
Mast Lighting Structures (HMLS) in Taiwan are investigated to obtain the WRD criteria and procedures
(Chien, et al., 2008a). The WRD criteria for HMS

598

Journal of the Chinese Institute of Engineers, Vol. 33, No. 4 (2010)

Table 1 Geometric properties for different types of HMS


Types of description

Diameter (m)

Thickness (mm)

Height (m)

slop (%)

/H (%)

High mast lighting structure


Signal support structure
Wind turbine tower
Telecommunication structure

0.5~1.0
0.5~1.0
2~4
0.5~1.0

6~8
6~10
6~15
6~10

30~40
15~20
50~100
30~60

1.0~1.5
0.0~1.0
1.5~2
1.0~1.5

< 15
< 15
<1
<1

Note: /H is deflection limit state as ref. Fig. 1

include two objectives. One is to define the identification of a yield criterion and the strength limitation
of the support structure. The other is to provide uninterrupted service by the tip utilities under extreme
wind. By using CFD (Computational Fluid Dynamics),
this study finds that the drag coefficients of an 8 or a
16-sided pole are close to those of a round cylinder.
This result comes from an analysis of eight different
sections of a cylinder for uniform and shear flows.
From the literature, we see that, usually, the
maximum response of a structure subjected to random loading cannot be evaluated in a deterministic
sense (Smith, 1988). Based on the gust factor
approach, simple semi-empirical closed form solutions that are amenable for practical use have been
developed for point-like and line-like structures
(Solaria, 1982). It is a useful method to determine
the across-wind aerodynamic analysis for WRD by
Scruton number in relation to a mass-damping parameter (Scruton, 1963). The wind flow around the bluff
body is characterized by a separation of the flow at
the leading edge corners. For bluff bodies with curved
surfaces such as the circular cylinder, the position of
the separation promotes the harmonic cross-wind
force variation on the structure. Therefore, the bluff
body may also enhance the vortex strength; the vortex shedding frequency may change to the frequency
of vibration. Through the feedback mechanism, the
frequency of the shedding of vortices can lock-in
to the frequency of motion of the bluff body. In wind
tunnel testing, it was found that large diameter round
monotubular mast arms would exhibit vortex shedding when there was no sign panel (Kevin, 1998). In
addition, when the tip had a sign panel installed, galloping occurred rather than vortex shedding. Indeed,
higher order mode effects are generally negligible in
wind engineering design (Kareem, 1981). However,
for HMS significant problems can occur in more than
one mode of a finite element program, particularly in
the first four modes (Wang, 1995).
In this paper, the contributions of high natural
frequency components are obtained by Eignvalue
analysis. The method to check vortex resonance and
galloping for higher order modes is also presented.
Because of turbulent winds in the atmosphere and
characteristics of the irregular bluff bodies of the

structures are complicated to deal with, a mathematical model, with interactive wind and structure is still
impossible at present.
This study includes four parts: (1) a survey of
geometric configurations and shape factors; (2) alongwind and across-wind response analysis; (3) develops criteria for WRD; (4) provides case application
for WRD procedures.
II. GEOMETRIC CONFIGURATIONS AND
SHAPE FACTORS
Wind force is the main consideration for the HMS
design, which is affected by the shape factor, projected
area, and angle of attack. Therefore, geometric configuration has an important influence on wind load.
Usually, this influence can be measured by means of
field measurement or wind tunnel test or CFD.
1. Geometric Configurations
HMS designs around different geometric configurations of tubular pole to meet requirements for
local code and clients specifications are shown in
Table 1. HMS has the characteristics of light weight
and cost efficiency. This is useful because freeways,
airports, ports, and stadium designs could not be effectively illuminated by conventional methods. To
meet specific mounting height requirements a shaft
may consist of more than one section to be assembled
on site by means of the slip on joint. Generally,
the geometric configurations of HMS are multi-sided,
slender and tapering. They are designed in the range
of 30 to 40(m) height with base diameters of approximately 460~1000 mm as shown in Fig. 1. Because
the diameter of the shaft is smaller than 1 m, it is
defined as a monotubular mast instead of as a monotubular tower structure.
2. Structural Properties
HMSes are usually mounted with some lamps
and lanterns on the tip of the structure. The combination of the structural slenderness with the concentrated masses at the top makes HMS fully aeroelastic
and unstable. Traditionally, if the cross section of an

599

C. W. Chien et al.: Wind-Resistant Design of High Mast Structures

Table 2 Drag coefficients on different sides of


regular polygon
Sides
(Shape)

Octagonal
N=8

Drag
coefficient

Hexagon
N=6

Cd

10

12

14

16

1.70 1.20 1.01 0.85 0.90 0.87 0.82

5D
Square
N=4

Circle
N=

Fig. 1 Cross-section and drift of the HMS

HMS is tapered, structural designers usually adopt


the average diameter or the diameter at z = 2/3 H (the
ratio of slope shall be less than 2%) (Wang, 1995). z
is the height above ground, and H is the height of the
structure. The recommended value of damping ratio
is 0.3~2.9% for steel mast and tower structures
(Smith, 1988). Slip jointed steel HMS structures have
high structural damping due to friction at the joints;
this energy absorbing mechanism will limit the resonant response to wind (Holms, 2001). Excluding fatigue design of vortex shedding-induced loads, the
damping ratio of a steel mast is assumed to be 1% in
this study (Wang, 1995).
3. Shape Factors
The geometric configuration of the structural element is taken into account through the use of a drag
coefficient, C d. The value depends on the geometric
configuration of the structural element and associates
with its Reynolds number. Reynolds number (Re = U
B/vK = 69000 U B) is the ratio of fluid inertia forces
in the flow to viscous forces, where U is the mean value
of the reference wind, B is the maximum width of the
structure, and v K (= 0.145 10 4m 2/sec) is the kinematic viscosity of air. For bluff bodies with curved
surfaces such as the circular cylinder, the separation
points are dependent on Reynolds number. For
. . 2
Bernoullis equation 1
2 a U + p = const, where the
constant holds along a streamline, and a is the air
density. Its usual to measure all pressures on the
structural surface of the far upstream side or the freestream wind at some distance from the structure. The
non-dimensional pressure coefficients C p is defined
as
p p0
(1)
Cp =
,
1 U2
a
2

8D

5D
20D

Fig. 2 Domain decomposition

where p-p0 represents the pressure difference between


local pressure p and far upstream pressure p 0 (Simiu
and Scanlan, 1996). Normally, the drag force, lift
force, and moment are all greatly affected by the
shape of the body; the Reynolds number (Re) is taken
into account especially for a circular shaft of an HMS.
A two dimensional CFD (Computational Fluid
Dynamics) analysis is always used in this study to
investigate the influence of shape factors. The simulations adopt the weakly-compressible-flow method
(Song and Yuan, 1988) together with a dynamic
subgrid-scale turbulence model (Germano et al.,
1991). In order to obtain the structures modal mesh;
the domain decomposition of sub-volumes is applied
as shown in Fig. 2. Assuming the approach flow is
uniform, the Reynolds number is 10 5 for all shapes.
Fig. 3 illustrates the contour of time average pressure in regular polygons with different numbers of
sides. The drag coefficients, obtained from the CFD
results, are shown in Table 2. For an HMS with a
polygon cross section, it may be assumed that C w = 0
for all cases, such that the total drag coefficient: Cd =
C w + Cl, Cw, C l are the values of the positive pressure
coefficient on the windward face and the negative
pressure coefficient on the leeward face, respectively.
Fig. 3 shows that the integration of the pressures over
the body surface results in a net force and a moment.
It shows that the integral value of the summation is
close to zero from the CFD results for a 16-side
section. Increasing the number of sides of the polygon C d leads to more and more regularity while C w
tends to be zero as shown in Fig. 4. Consequently
for engineering purposes, it is acceptable that C w = 0

600

Journal of the Chinese Institute of Engineers, Vol. 33, No. 4 (2010)

(a) 4 Sides
PA:

-2 -1.6 -1.2 -0.8 -0.4

(b) 6 Sides
0

0.4 0.8

PA:

(c) 8 Sides
PA:

-2 -1.6 -1.2 -0.8 -0.4

-2 -1.6 -1.2 -0.8 -0.4

0.4 0.8

PA:

-2 -1.6 -1.2 -0.8 -0.4

0.4 0.8

-2 -1.6 -1.2 -0.8 -0.4

0.4 0.8

0.4 0.8

0.4 0.8

(f) 14 Sides
0

0.4 0.8

PA:

(g) 16 Sides
PA:

(d) 10 Sides

(e) 12 Sides
PA:

-2 -1.6 -1.2 -0.8 -0.4

-2 -1.6 -1.2 -0.8 -0.4

(h) circular
0

0.4 0.8

PA:

-2 -1.6 -1.2 -0.8 -0.4

Fig. 3 Contour of average pressure

is used for regular polygons with more than 16 sides.


Usually, the HMS takes rounded corners into account,
so, it seems reasonable to assume C w = 0 as above.
Table 3 shows the comparison of drag coefficients of
regular polygon in different codes.

III. ALONG-WIND RESPONSE


The along-wind responses, such as maximum
force, displacement, root mean square (R.M.S.), and
peak acceleration, are usually obtained by the gust

601

C. W. Chien et al.: Wind-Resistant Design of High Mast Structures

Table 3 Drag coefficients of regular polygon in different building codes


Cross section

H/B

Wind direction

Square
Hexagonal or octagonal

25

Wind normal to face

1.4

1.3

Wind along diagonal

1.5

1.1

1.0

Wind normal to face

1.4

1.2

1.0

Round
ASCE7-05 D q Z > 5.3 ;

Rough
(= 0.02D)

Wind normal to face

0.9

0.8

0.7

D in m, q Z in N/m 2
Taiwan-06 D q Z > 1.5;

Very rough
(= 0.08D)

Wind normal to face

1.2

1.0

0.8

D in m, q Z in kgf/m 2
China code-01 w 0D 2 0.015 w 0
in N/m 2

Moderately
Smooth

0.6

0.5

0.45

0.7

0.6

0.5

1.2

0.8

0.7

Wind China
Normal Taiwan
to face ASCE7-05

Round
ASCE7-05 (D q Z 5.3);
D in m, q Z in N/m 2
Taiwan-06 D q Z 1.5;
D in m, q Z in kgf/m 2

Wind normal to face

2.5

2
Cl
Cw
Cd

1.5

0.5

0.5

Cd

1.5
Cl, Cw

g = 2 ln( T0) +

,
2 ln( T0)

where g is the peak factor for a Gaussian process


(Davenport, 1964). It depends on the time interval
for which the maximum value is calculated, and the
frequency range of the response. = 0.577 is Eulers
constant; T 0 is the period when the peak response is
assumed to occur, and is the expected frequency.
In the narrow-band (or resonant response region),
is equal to the natural frequency n, T 0 is the 1 hr. =
3600 second, and the resonant peak factor g R from
Eq. (2) may be evaluated as follows

-0.5

g R = 2 ln(3600n) +
-0.5

(2)

8
12
Sides of polygon

16

20

-1

Fig. 4 Drag coefficients on different sides of polygon

effect factor method. This approach is based on the


maximum responses that can be represented by the
corresponding mean response multiplied by a gust
effect factor. Usually, the gust component is considered as a random process and may be described by a
gust spectrum with variance x2.
1. Maximum Response
Davenport (1964) derived a gust factor such that
the root mean square component would be exceeded
with a probability of 50%, and recommended the following expression for this peak factor

0.577
.
2 ln(3600n)

(3)

In the broadband region (or non-resonant


region), one may evaluate a non- resonant peak fac~ 3.5. In the basic or initial design stage, one
tor g B =
can use the resonant peak factor g R to replace the g
peak factor, then use the gust effect factor to design.
However, using normal spectral methods and neglecting second order terms, one may combine the gust
spectrum with a transfer function for forces described
by its variance x2. Furthermore, using these factors,
the maximum total load effect EMAX may be obtained

by summing the mean load effect E together with the


factored broad-band and narrow-band load effects as
follows (Smith, 1988):
E MAX = E +

[g B B(E)] 2 + [g R R(E)] 2 ,

(4)

where B(E) is the R.M.S. value of the non-resonant

602

Journal of the Chinese Institute of Engineers, Vol. 33, No. 4 (2010)

load effect, and R(E) is the R.M.S. value of the resonant load effect. Therefore, the expected maximum

value of random response X is obtained by X MAX = X

+ g . x. X is the mean value of random response. x


is the gust spectrum with standard deviation. As the
purpose of the simplification is to calculate the formula of a peak factor (g) for an engineering design,
Solari (1982) proposed a design peak factor g D to replace g:

g D = 1.175 + 2ln n T0 .

F(z) = q . Cd . B(z) = 1 . a . V 2(z) . Cd . B(z), (6)


2
where q is the velocity pressure, B(z) is the width of
the structure at height z. Furthermore, we assumed
(z) = ( z ) ; (z) is the fundamental model shape at
H
height z. is the mode exponent and can be derived
by integrating the mean intensity of wind loading over
the height of HMS. The expression at any point is
given by

V (z) 2 B(z) (z)dz ,

(7)

where F d is the total drag force, therefore, the mean


displacement of mean along-wind velocity at the average height point is obtained by

Xave = d ; K 1 = m 1(2 n 1) 2 ,
K1

(9)

The second-order term of u 2(z, t) can be ignored

in comparison with V 2(z), hence, the fluctuating component will be given by

The wind load calculation may be divided into


three parts; including along-wind load, across-wind
load, and torsional load. For most tall symmetrical
buildings, only the fundamental modes in alongwind
and acrosswind directions will be considered in the
structural design. In this study, we assumed the structure
will undergo aerodynamic actions partly distributed
along the axis of the shaft and partly concentrated in
the geometrical center of the masses; therefore torsional load can be ignored. (Solaria, 1999) Alongwind load on a structure is dependent on the structures
properties, shape, height, and surroundings. Commonly,

we may separate the gust wind velocity V(z) + u(z, t)

into two parts, a mean V(z) and a turbulence compo


nent u(z, t). The mean component V(z) is measured
every 10 minutes by means of wind velocity and the
turbulence component u(z, t) is calculated at average
intervals of 10 sec. The mean intensity of wind loading at any point on an HMS can be calculated by the
following equation:

Fd = 1 C d a
2

F(z) = 1 a[V(z) + u(z, t)] 2 . C d . B(z).


2

(5)

2. Maximum Along-Wind Load

for the first mode of the structure, respectively. The

wind speed vector at any point V(z, t) = V(z) + u(z, t)


may be expressed as the sum of the mean wind speed
vector and a dynamic component. The distributed load

ing at height z is F(z) = F(z) + f (z, t), where F(z) and


f (z, t) are the mean and fluctuating components,
respectively. If expressed in terms of velocity, Eq.
(6) becomes

(8)

where K1 is the stiffness for the first mode of the structure.


m1 and n1 are generalized mass and natural frequency

f (z, t) = a . V(z) . u(z, t) . C d . B(z).

(10)

Hence, the load effect E(t) on the structure element will be

dE(t) = a . V(z) . u(z, t) . C d . B(z) . dz.

(11)

To analyze the dynamic response of the


structure; the randomly distributed fluctuating forces
are considered as discrete forces.

F i(z, t) = F(z) + f (z, t)

(12)

Hence, the total wind load on the structure will


be
N

F = F1 + F2 +

+ Fi

+ FN = Fi .

(13)

The square of the total load will therefore be


F 2 = F12 + F22 +

N N

+ FiF j

+ FN2 = FiF j . (14)


i

The mean load square can be obtained by averaging Eq. (14) over an appropriate period, e.g. one
hour. Using Eq. (4), the maximum total load effect
may be obtained by summing the mean along-load
effect, the factored broad-band and narrow-band load
effects:

F max = F + [(g B BF) 2 + (g R RF) 2] 1/2,

(15)

where F is the value of the mean along-load; BF is


the R.M.S. value of the non-resonant load, and RF is
the R.M.S. value of the resonant load.
3. Maximum Along-Wind Displacement
The total along-wind displacement X max(H) can
be divided into two parts: First, the average displace
ment x(H)
caused by average wind speed. Second,
the turbulent displacement x max (H) caused by the

603

C. W. Chien et al.: Wind-Resistant Design of High Mast Structures

turbulent wind speed. Hence, the total along-wind


displacement of a structure subjected to along-wind
force is given by

.
Lock-in region =. Ur ~ 1.3 Ur
Frequency
Resonance of
vortex

X max(H) = x (H) + [(g B Bx) 2 + (g R Rx) 2] 1/2


= x (H) + x max(H),
(16)

Vortex shedding frequency


Natural
frequency of

where BX is the R.M.S. value of the non-resonant


displacement; RX is the R.M.S. value of the resonant displacement.

Flow velocity

Fig. 5

IV. ACROSS-WIND RESPONSE


Across-wind response is more likely to arise
from vortex shedding or galloping. These motioninduced wind loads, rather than those for along-wind,
have become significantly important for HMS design.
Therefore, these across-wind response behaviors also
need to be investigated for the WRD criteria.
1. Vortex Shedding
Usually, when the mean design wind speed (or
tip wind speed V h) is greater than the critical wind
velocity of vortex resonance Ucr(z), the structure falls
in the resonance area. Although vortex shedding can
lock-in and continue as the wind velocity increases
or decreases slightly, if the wind velocity changes by
more than 30%, the vortex shedding will stop. The
result is a considerable broadening of the region of
induced vibration. Consequently, its different from
the narrowband mechanical amplification function as
shown in Fig. 5. However, V h should be stated by
the local code or one can use the following equation
(GB50009, 2001):
V h = 40 z(H) ro ,

(17)

where o is basic wind pressure, r is equal to 1 in a


return period of 50 years, Z is the coefficient of the
pressure exposure at z height. The critical wind velocity of vortex resonance at height z, U r(z), can be
obtained by Eqs. (18) and (19):
f B(z)
St = s
,
U r(z)

5 B(z)
U r(z) =
,
Tj

Evolution of vortex-shedding frequency with wind velocity over elastic structure

the first to fourth modes shall be considered in the


design of HMS when the U r(z) is between 0.50 and
1.30 V(z cr ); V(z cr) is the mean design wind speed at
zcr, z cr = 5/6 height of the structure. Furthermore, the
equivalent static pressure range to be used for fatigue
design of vortex shedding-induced loads shall be the
following (AASHTO, 2001)

PVS =

0.613 U r(z) 2 C d I F
,
2

2. Galloping
Galloping is a typical instability of slender
structures, having special cross-sectional shapes such
as rectangular or nonsymmetrical cross-sections.
However, galloping is different from vortex shedding,
which results in large-amplitude, resonant oscillation
in a plane normal to the direction of wind flow. When
the structural damping is equal to the aerodynamic
damping, the critical wind velocity of galloping (v oc)
can be obtained from the following equation (Wang,
1995):
2 CV
(0)
a DL

(18)

(19)

where T j is the natural period of the structure ( j = 1,


2 ... n). Across-wind loads due to vortex shedding in

(20)

where I F is fatigue factor; is the damping ratio for


vortex shedding, which is conservatively estimated
as 0.005. The damping ratio of the structure will decrease with vortex shedding-induced for these higher
order modes.

v0c =

where f s is the vortex shedding frequency, B(z) is the


diameter at height z, and the Strouhal Number S t of
the round cylinder is adopted as 0.2. Then, Eq. (18)
can be transformed to Eq. (19):

Mechanical
resonance

H
0

(21)

Z B(z) 2(z)dz

where CV = 2 . M * . . is the generalized damping,


and is the damping ratio of the structure. M * =
H
m(z) . 2(z)dz and are the generalized mass and
0

the natural circular frequency, respectively. m(z) is


the mass per unit length along the structure. is the
angle of wind in the main x-direction, when = 0,
D L(0) is the derived value of the coefficient of galloping force DL( ), and the Eq. (22) is

604

Journal of the Chinese Institute of Engineers, Vol. 33, No. 4 (2010)

Table 4 Coefficient of drag derivatives (Wang,


1995)
Cross section Re
1

DL(0) Cross section

66000 2.7
1

33000 3.0

Re

Construct finite
element model

DL(0)

Solve gust effect factor


G = ? (GEF/ESWL)

2000~ 10.0
20000
75000 0.66

Apply dynamic
analysis
Construct wind
loading model

sin
DL( ) = [ D( )
+ L( ) 1 ] , (22)
cos
cos 2

where D( ) is a coefficient of along-wind force or


is defined as the coefficient of drag force. It is equal
to the front s1 with the absolute back side s2 as given
by D ( ) = s1 + | s2 |; L ( ) is the coefficient of
across-wind force. D( ) and L( ) are obtained by
wind tunnel testing, D L is shown in Table 4. By conducting wind tunnel tests, an equivalent static vertical shear of 1000Pa was determined for the galloping
phenomena (AASHTO, 2001). The magnitude of this
vertical shear pressure range shall be equal to the
following: P G = 1000 I F (Pa) where I F is fatigue
factor. For example, if a 3 m by 1 m sign panel is
mounted to a horizontal mast arm, a static force of
3000 IF, Ng should be applied vertically to the structure at the center of gravity of the sign panel.

Check vortex
galloping
stability
displacment

Apply aeroelatic
analysis

(a)

G=?

Tn

< 1 sec.

> 1 sec.

Rigid-Str.

WindSensitive str.

G = 2.33

ASCE7Code/ESWL

V. WIND-RESISTANT DESIGN CRITERIA


Generally speaking, there are two objectives for
the WRD criteria. First, an HMS should provide uninterrupted service under typhoons or strong wind
strikes. Second, an HMS should be able to avoid
collapse and avoid structural damage from missile
impact induced by typhoons. Usually, some serviceability requirements of HMS are strict, maintaining
the function of utilities, such as telecommunications,
microwaves, and wind-power turbines. However, an
HMS is especially sensitive to across-wind vibrations
caused by vortex shedding or galloping, which will
lead to wind-excited instability. Therefore, designers should check the critical wind speed of vortex
shedding and galloping. To meet the requirements
for WRD of an HMS, some criteria and procedures
have been outlined and presented in the flowcharts
as shown in Figs. 6(a)~(b). These desiderata are stated
as follows:
1. Design Parameters Requirements
A good geometric configuration is important for
WRD design to obtain good shape factors. Commonly,
the parameters of wind loading calculation are obtained from the local building code as well as from
clients specification.

(b)

Fig. 6

(a) Wind-resistant design procedure for HMS (b) procedure of gust effect factor calculation

2. Buckling Resistance and Serviceability Requirements


If the D/t ratio of HMS is larger than 125, the
WRD should counter local buckling of the shaft, and
control the ratio of D to avoid the buckling of the
t
shaft from wind loading, which is given by the following equation (Schilling, 1965):
D < 0.1259 E s ,
t
Fy

(23)

where D is the diameter of 2/3H height, t is thickness


of shaft, E s is modulus of elasticity, and F y is yield
stress. Further, we may control the buckling failure

C. W. Chien et al.: Wind-Resistant Design of High Mast Structures

605

by limiting the ratio of lateral deflection-height ( /


H) as shown in Fig. 1. Usually, there are many limitations for lateral deflection as specified in different
loading and design codes. In some cases, these limitations are given as design recommendations instead
of requirements. There are two main reasons for
adopting the wind drift deflection limitation in WRD
(a) to reduce the effects of motion perceptibility; (b)
to limit the P-Delta or secondary loading effects
(Mendisv and Cheung, 2007).
Table 1 depicts the horizontal deflection limitations for vertical support structures, such as HMS,
which should be designed to meet the following
criteria: (a) under group-I loading (dead load only),
the deflection at the tops of vertical supports with
transverse load applied shall be limited to 2.5% of
the structure height; and (b) under group-II loading
(dead load plus wind), the deflection shall be limited
to no more than 15% of the structure height.

mode, it can be regarded as a single degree of freedom in dynamic analysis. If the HMS is an important and flexible structure, its suggested one perform
wind time history analysis. The procedure in the time
history method for applying the wind load model contains two distinct parts. First, the wind velocity time
series for a specific height must be determined. This
part includes the use of an approximation of the linear wind velocity power spectrum, which is essentially the energy density and is a function of frequency.
Second, the wind velocity time series is then extruded
along the height of the structure following ASCE-705 and thus producing a wind velocity profile. Using this wind velocity profile, the loading due to wind
velocity is determined based on drag force calculation,
or specifically a relative force equation known as
Morisons equation.

3. Static Analysis

Typically, wind load in equivalent static wind


loads can be derived by two methods. First, one can
calculate the mean static wind pressure and dynamic
wind pressure. From Eq. (16), the gust effect factor
of G is defined as

The Finite Element Method is used to determine


the static wind loading. Based on the results of Table
2, this study suggests that one may use Eq. (24) for
regression of C d:
C d = 0.6 + 1.44 ,
N

(24)

where N is the number of sides of the HMS; for circular shape, N is equal to infinity as shown in Fig. 1.
The recommended value of shape factor (drag coefficients of the shaft) is 0.82 for sixteen-sided polygons or circular section of HMS as per AASHTO code.
4. Dynamic Analysis
The application of the geometric stiffness matrix is a general approach to include these effects during
the analysis of all types of structural systems (Wilson
and Habibullah, 1987). In this study, an iterative
approach was utilized for dynamic response spectrum
analyses with consideration of second-order moment
effects. It is difficult to calculate the natural period
of the HMS. Therefore, it is necessary, by means of
the finite element method (FEM), to obtain the natural period of the structure correctly. In order to obtain the natural period for higher order modes of the
HMS and to check the critical wind velocity of vortex resonance in across-wind force, the Eignvalue
Method is used. Generally, natural period can be used
to determine the rigidity or flexibility of an HMS. If
the equipment weight at the tip is much greater than
the body weight of the pole, it can be treated as a
point structure; otherwise, it should be regarded as a
multiple mass structure. If it is governed by the first

5. Gust Effect Factor (GEF) Method

x (H)
G = 1 + max
x(H) .

(25)

Second, one can multiply mean wind pressure by


a gust effect factor of G. Substituting Eq. (25) into
Eq. (16), one can obtain the maximum displacement
Xmax(H) = G . x(H), which is the maximum fluctuating
displacement of mean wind velocity multiplied by the
gust effect factor of G value at the average of z height.
By the way, Fig. 6(b) is also another gust effect factor
method procedure to obtain the gust effect factor of G.
It is a good identification, such that the natural period
of one second can be used in the discrimination of the
rigidity and flexibility of a structure (Chien and Jang,
2008a). If the natural period is larger than one second,
structural designers will need to perform a more rigorous analysis as defined in the ASCE7-05 code, which
is shown in Fig. 6(b). If designer want to perform a
more rigorous analysis as shown in Fig. 7, or follow
criterion 4 to perform a wind time history analysis.
6. Aerodynamic Analysis Criterion
Aerodynamic analysis should be considered in
WRD due to the wind-sensitivity of HMS structures.
We suggest that Eq. (26) can be used as a criterion to
decide whether to perform aerodynamic analysis for
an HMS or not (Melbourne, 1997).

S c = 4 K s = s < 5 ,
a

(26)

606

Journal of the Chinese Institute of Engineers, Vol. 33, No. 4 (2010)

Wind
velocity/force/response
Mean wind
velocity

Uwind

UHMS

Urel

Aerodynamic
admittance

Wind loading
spectrum

Mean wind
force

Background
response

Wind gust
response

Turbulent wind
velocity

Mechanical
admittance

Resonance
response

Fig. 8 Relative velocity of wind and HMS

E.S.W.L.

Fig. 7 Procedure of equivalent wind static loading calculation

where, the mass-damping parameter K s = m / aD 2,


m is the mass of the structure. a =0.125 kg-s 2/m 4,
and s is the mass density of the structure. When the
Scruton number S c is less than 5, one also needs to
perform aerodynamic analysis (e.g. wind tunnel test)
or aeroelastic analysis to obtain more accurately parameters and results.

Find the natural


period of
structure

Find V(zcr)
basic design
wind speed as local code
Non-tapered calculation
vortex shedding fatigue
force
as Eq. (20)

Taper/
Non-tapered str.

Find the critical wind


speed as Eq. (19)
Check
Re

Check
V(zcr) > Ur(z)

Re < 5 105

7. Aeroelastic Analysis Criterion


Re 3.5 106

If the HMS moves, the relative motion of the


interaction of wind load and structure, generates an
aeroelastic force. Morisons Eq. can be used to determine the force from the given wind velocity as
shown in Eq. (27) (John, et al., 2006):
F = 1 a AC d U wind U wind ,
2

(27)

where A is the tributary projected area for the nodal


point, U wind is the wind velocity at the nodal point.
For the HMS analysis herein referred to as the pole,
the movement of each point on the pole must be considered using this wind velocity (U HMS ). The real
forcing function must be computed from the relative
velocity (Urel) between the pole and the wind as shown
in Fig. 8.
8. Across-wind Response Analysis Requirements
One sees, from Fig. 9 that the across-wind load
does not need to be considered, when the U r(z) falls
outside the range of 0.50 and 1.30 V(z) cr as defined
here. Where V(z) cr = Vh is set to be the mean design
wind speed (m/sec), it can be obtained from local
code. To obtain the U r (z), Eq. (19) can be used to

Calculate the acrosswind force as Eq. (28)

Fig. 9 Procedure for across-wind force calculation

determine whether or not to perform across-wind response analysis by vortex shedding. Further, the
across-wind force can be derived by using Eq. (28)
(Tsai, et al., 2006):
Wrz = U r(z) 2 Z C r A ,
h

(28)

where W rz is the across-wind force at height z, C r is


the wind force of vortex resonance. Furthermore Eq.
(29) can be used to obtain the tip displacement of HMS
in the across-wind response analysis (GB50009, 2001):
Xc =

U r(z) 2 D (Z)
,
8000 m 02

(29)

Furthermore, the total wind response or force


X total can be obtained by the following Eq. (30).
X total =

X a2 + X c2 ,

(30)

607

C. W. Chien et al.: Wind-Resistant Design of High Mast Structures

Table 5 Variables for HMS case


Section no.
Shaft mass (kg)
Luminaries (kg)
Height (mm)
Laps length (mm)
Laps length (mm)
Total length (mm)
Thickness (mm)
Base diameter (mm)
Diameter/Thickness

(1)

(2)

(3)

Total = (1) + (2) + (3)

1804

984

596

9400
1400

9800
1000
1400
10800
6
766
127

10800

3384
1000
30000
upper joint
lower joint

10800
8
1000
125

where X a and X c are along-wind and across-wind response (e.g. force or displacement), respectively. The
direction of resultant response (e.g. force or displaceX
ment) has an angle of = tan 1 a .
Xc
Figure 10 illustrates the check procedure for
galloping effect. Eq. (21) is used to obtain the critical wind velocity for galloping. When it is less than
the mean design wind velocity (or tip wind velocity),
one should calculate the galloping force. Generally,
the galloping force is derived from wind tunnel tests
or the following formula: P G = 1000 . I F for HMS
(AASHTO, 2001).
9. Stability Analysis Requirements
An HMS shall be considered to need guy cables,
or extra weight and foundation at the base for stability,
when the overturning moment (Mo) is greater than the
resistant moment (M r ). The safety ratio factor for
M
overturning stability is defined as FS = r 1.5. For
Mo
across-wind response analysis, when the total tip displacement of HMS is greater than the serviceability
limitation of the structure, its necessary to check the
FS ratio for the HMS stability. Furthermore, the criteria formulation of stability may use the plastic design criterion, which is shown in Eq. (31) (ASD, 1989):
Cm M
P +
1,
Pcr (1 P ) M
cr
PE

(31)

where P E is Euler critical loading; P is applied axial


compression load; Pcr is nominal strength for the axially loaded compression member, M is applied primary bending moment, M cr is nominal moment
strength in the absence of axial load; C m is amplification coefficient, C m is equal to 1 for the equal moment force of a cantilever pole; when a concentrated
force is applied to the cantilever at the tip of the cantilever pole, it is equal to 0.78. For uniform horizontal force on a cantilever pole, it is equal to 0.49.

1000
10800
6
522
87

Define structure
properties
General damping
ratio / mass

Define
Basic wind pressure
return period

Calculate
tip of wind
velocity/Critical
wind velocity

Define geometry
properties
Height/Width of
structure or
equipments

Define
coefficient of drag
derivatives / height

Wind tunnel test


/Calculating
PG = 1000 . IF

Fig. 10 Procedure for galloping check

The above criteria all have significance in HMS


analysis and in the design process. The first five relate to strength and serviceability and the last relates
to stability considerations.
VI. DESIGN PROCEDURE AND CASE STUDY
The previously described procedures and criteria have been applied to WRD of HMS. In order to
quantify the implications of the above results, an example is presented and described below. The
monotublar pole under investigation is the prototype
of an HMS adopted by Kaohsiung Mass Rapid Transport in Kaohsiung (KMRT), Taiwan. In this case,
the monotublar pole is a 30 m high mast manufactured from pressed steel sheet welded longitudinally.
It is a sixteen sided polygon, continuously tapering,
with a hollow body. Table 5 summarizes the main
geometric properties. There is a set of luminaries
(weight = 1000 kg), and the project area is about 2.0
m2 at the summit. The shaft consists of three sections,
assembled by slip on joints as shown in Fig. 11. The
design procedure and a WRD case are presented to
illustrate and identify the calculation of the design

608

Journal of the Chinese Institute of Engineers, Vol. 33, No. 4 (2010)

530

522* 240*6*10800

144

venus1000
161

510

22.5 22.5

3000

766*484*6*10800

1000

45

1120

1000

940

25

1800

1000* 718*8*10800

1230 TH.30

N.32 ANCHOR BOLTS ASTM A36IS275JRIM36 1

Fig. 11 High mast lighting structure

criteria for HMS. In this case we also compared different design criteria and requirements from Taiwan,
China, and AASHTO codes.
1. Design Procedure
Step 1: Design parameters for wind loading
Usually, the design wind speed of WRD should
be determined in accordance with local code and to
meet the clients requirements. Therefore, for the client,
design wind velocity tolerance is 72 m/s. In order to
investigate this case for variables for wind the HMS
parameters are shown in Table 6. In this case study,
basic wind pressure is o = 0.8 KN/m2; r = 1 in a return period of 50 years, the factor of the pressure exposure z = 0.62, the coefficient of the exponent in the
power law = 0.22 (GB50009, 2001), the drag coefficient is assumed to be 0.6 for a shaft (Tsai, et al., 2006);
a shape factor of 1.0 was assumed for luminaries or
blades of structures at the top (John, et al., 2006).

Step 2: Buckling resistance analysis


First, one shall calculate the maximum ratio of
D/t (= 127), it is larger than 125 as shown in Table 5.
Second, in order to avoid the local buckling of the shaft,
the maximum ratio of shaft should be controlled to
6
less than Eq. (23) (i.e. D = 127 0.1259 2.1 10
t
3500
= 717.9). Furthermore, in order to avoid instability
of the mainframe, one should control the maximum
deflection ( /H) to meet the requirement of serviceability limit of HMS as shown in Table 1. In this
case the limit is if the /H ratio is larger than 15%
(with the consideration of wind and gravity load as
Step 7) for HMS. Then, the stability analysis in Step
8 is needed to check the stability.
Step 3: Dynamic analysis
In this work, SAP2000 was employed, 15 modes
were utilized to include at least 90% of the participating mass for principal horizontal directions, and

609

C. W. Chien et al.: Wind-Resistant Design of High Mast Structures

Table 6 Parameters used for wind force and structures response calculation
Parameter

Decription

Taiwan

unit

Zd
Z0

U 10
u*

W
Tn
Es
Fy

zero plan displacment


roughness length
coefficient of the exponent in the power law
basic design speed
shear velocity
damping ratio
weight of structure
natural period
elastic modulus
yield strength

5.00
0.30
0.25
37.5
5.33
0.01
4384
1.54
21000000.00
3500

m
m
m/s
m/s
kg
sec
ton/m 2
kg/cm 2

Table 7 Periods and effective modal mass ratio of the HMS case
Mode period
No.
1
2
3
4
5
6

Individual mode (%)

(sec)
1.5412
1.5412
0.2456
0.2456
0.0734
0.0734

UX
19.80
44.42
23.24
0.12
3.48
6.52

UY
44.42
19.80
0.12
23.24
6.52
3.48

Cumulative sum. (%)


UZ
0.00
0.00
0.00
0.00
0.00
0.00

UX
19.80
64.22
87.47
87.59
91.07
97.59

UY
44.42
64.22
64.35
87.59
94.10
97.59

(a) 1st mode


Tn = 1.541 (sec)
-175.

(b) 3rd Conic section


t = (6 mm)

-140.

-105.

-70.

(v) 2nd Conic section


t = (6 mm)
-35.

0.

UZ
0.00
0.00
0.00
0.00
0.00
0.00

(d) 1st Conic section


t = (8 mm)
35.

70.

105.

SAP2000-Resultant Moment MAX. Diagram (group-II)-kgf-cm Units


Fig. 12 30 m HMS mode shape and modal section

the first six dominant modes are presented in Table


7. The natural period of the structure (T n) was obtained by eignvalue analysis. The results indicate that
the first two fundamental modes of the model are
dominated by lateral translations. Since the HMS case

is a symmetrical structure, their vibration periods are


close to each other. The modal section and first mode
shape of the structure are shown in Fig. 12(a). One
checks whether the model of dynamic analysis is
correct, to meet the requirements of the vibration

610

Journal of the Chinese Institute of Engineers, Vol. 33, No. 4 (2010)

Table 8 Segments of wind gust factor (Taiwan code-2006)


NODE

B(z)

IZ

V(Z)

Rn

RB

RL

1
2
3
4
5
6

30.0
24.0
19.9
18.4
11.4
9.95

0.49
0.49
0.73
0.73
0.73
1.00

0.27
0.27
0.27
0.27
0.27
0.27

72.0
46.67
44.5
43.7
38.7
37.45

0.89
0.89
0.89
0.89
0.89
0.89

0.23
0.21
0.20
0.20
0.19
0.19

0.96
0.94
0.94
0.94
0.93
0.93

0.94
0.91
0.86
0.86
0.85
0.79

2.10
1.68
1.62
1.60
1.48
1.43

3.91
3.35
3.27
3.24
3.09
3.03

Table 9 Calculation of along-wind force by different codes (V h = 72 m/sec)


H(m)

Taiwan-1996 (kg)

AASHTO-2001 (kg)

China-2002 (kg)

Taiwan-2006 (kg)

30.00
24.00
19.90
18.40
11.40
9.95

1999
1668
584
2679
689
4593

1836
1674
613
2859
816
5597

3558
2179
1061
1497
986
2218

2147
1704
608
2820
770
5192

Total

12212

13394

11499

13241

25

shown in Table 8. Consequently, characteristic samples


of the results obtained from the static analysis of the
entire HMS are as shown in Figs. 12(b) ~ (d). Alongwind force is 13241 kg obtained by following Taiwans
2006 code, and its close to the result 13394 kg from
the AASHTO 2001 code as shown in Table 9.

Node-location (m)

20

15

Step 5: Aerodynamic analysis


Mass density s is equal to 322.35 ( s = 4384/
13.6 = 322.35), S c = s / a = 0.01 322.35/0.125
= 25.79. However, the value of S c = 25.79 is greater
than 5.0; this result can be neglected for aerodynamic
analysis in this example.

10

Comparison of shapes
1st Mode
2nd Mode
3rd Mode

Step 6: Across-wind response analysis


0

-3 -2.5 -2

-1.5 -1 -0.5

0.5

1.5

2.5

Displacement (cm)
Fig. 13 Comparison of mode shapes

theory. Fig. 13 shows the comparison mode shapes


of HMS. In this case, pretty good agreement for the
first three modes was observed.
Step 4: Static analysis and GEF analysis
Static wind load analysis is under step 1 to obtain the design wind pressure tolerance for HMS.
Along-wind loads and gravity loads are considered
in static analysis. Furthermore, calculation of gust
effect factor G is based on Taiwans 2006 code as

Equation (19) is used to obtain the Ur(z) in acrosswind. Comparison of the design wind speed tolerance and tip wind speed V h to check the vortex
shedding resonance is shown in Table 10. However,
when U r(z) < 1.3 V h one should consider the acrosswind force, as shown in Table 11. The across-wind
force is obtained by following Taiwans 2006 code.
It is about 40% of along-wind force. Finally, maximum along-wind displacement (Xa) is obtained by Eq.
(16) as shown in Table 12. Furthermore, one can solve
the tip displacement in across-wind by Eq. (29), as
follows:
2
2
7
X c = 15.45 20.246 0.76 0.522 10 = 4.15 cm ,
4 8000 0.01 4384

where = 0.01 is the damping ratio of structure; the

611

C. W. Chien et al.: Wind-Resistant Design of High Mast Structures

Table 10 Calculation of across-wind resonance fatigue forces


i-mode
(i = 1~6)

Tj
(sec)

U r(z)
(m/sec)

Vh
(m/sec)

Across-Wind
resonance

1-mode
2-mode
3-mode
4-mode
5-mode
6-mode

1.541
1.541
0.246
0.246
0.073
0.073

2.47
2.47
15.45
15.45
52.05
52.05

72.00
72.00
72.00
72.00
72.00
72.00

U r(z) < 20 m/s


U r(z) < 20 m/s
Lock-in
Lock-in
U r(z) > 20 m/s
U r(z) > 20 m/s

Across-Wind
fatigue force (kg)
37
37
1462
1462
16610
16610

Table 11 Calculation of across-wind forces


Node

Z
(m)

B(z)
(m)

U r(z)
(m/sec)

A
(m 2)

W
(kg)

Cr

Across-Wind
force(kg)

1
2
3
4
5
6

30.00
24.00
19.90
18.40
11.40
9.95

0.49
0.49
0.73
0.73
0.73
1.00

9.92
9.92
14.76
14.76
14.76
20.33

1.46
2.46
2.03
3.09
3.07
0.73

69.04
116.22
95.87
145.51
144.66
34.19

4.03
4.03
4.03
4.03
4.03
4.03

580.51
781.75
1183.39
1660.83
1022.94
400.38

5629.81

Table 12 Calculation of displacement and gust factor of HMS case


Description

Tip of velocity
y

Symbols
By equation
Unit
Taiwan
China code
KMRT

Vh
Eq. (17)
m/s
49.35
65.40
72

Along-Wind
d
average

X ave
Eq. (8)
mm
115.94
203.59
246.75

Along-Wind
d
maximum

Across-Wind
displacement

Total of
displacement

Xa
Eq. (16)
mm
154.25
259.98
307.75

Xc
Eq. (29)
mm
41.5
41.5
41.5

X total
Eq. (30)
mm
159.73
262.83
310.11

Gust factor
G
Eq. (25)
2.33
2.28
2.25

Table 13 Calculation of critical velocity of galloping for HMS case


Description
Height
Width of lighting equipment
Coefficient of the pressure exposure at height z
Coefficient of return period
Design wind pressure
Natural period
Generalized mass
Coefficient of drag derivatives
Generalized damping
Tip of Velocity
critical wind velocity of galloping

U r(z) is equal to 15.45 (m/sec) in the second mode.


D is equal to 0.76 meters. Therefore, the total displacement response (X total ) can be obtained by Eq.
(30), and the gust factor G equal to 2.33 can be calculated by Eq. (25) as shown in Table 12, this value

Symbols

Data

Unit

H
B
z
r
d
Tj
M*
DL(0)
Cv
Vh

30
2.5
1.42
1.1
3.24
1.54
4.479
3
0.3653
62.54

m
m

v ov

65.37

KN/m 2
sec
ton
ton/s
m/s
m/s

of G is different from Taiwans 2006 code as shown


in Table 8.
The critical wind speed of galloping is checked
by Eqs. (21)~(22) as shown in Table 13. Its equal to
65.37 m/s and larger than the design wind speed

612

Journal of the Chinese Institute of Engineers, Vol. 33, No. 4 (2010)

Table 14 Calculation of stability ratio for HMS case


Symbols

Data

Unit

E: Elastic modulus
F y : Yield stress
T: Thickness of shaft
A: Area
P E: Euler critical loading = 2 . EI /(K . L) 2
P: Axial compression load
C m: Constant
M: Primary bending moment
M cr: Nominal moment strength in the absence of axial load

2.04E + 06
3550.00
0.80
118.88
74300.04
8340.00
1.00
2.12E + 06
3178589.60

kg/cm 2
kg/cm 2
cm
cm 2
kg
kg

Stability ratio

0.88

tolerance 62.54m/s. Therefore, this case is not susceptible to galloping-induced vibration.

Under the expected design wind speed 72m/s,


the maximum horizontal displacement is equal to 140
cm for dead load plus wind load (i.e. D.L. + W.L.),
which is less than 15% of the structure height (AASHTO,
2001) as shown in Fig. 14. Second-order moment is
considered in the SAP2000 analysis for this case.

Height to Displacment

160

WL

140

Displacement (cm)

Step 7: Serviceability analysis

kg-cm
kg-cm

0.75 (WL + LL)

120
100
80
60
40
20
0

30

24

19.9
18.4
Height (m)

11.4

9.95

Fig. 14 Height to displacement

Step 8: Stability analysis


Stability analysis shall conform to the requirements
of the standard specifications in ASD and AASHTO.
Table 14 depicts the HMS safety factor, the ratio of
stability = 0.878 which is less than 1.0 in this study.
2. Discussion
The major observations of this study are summarized as follows:
(1) The WRD procedure presented in this paper for a
prototype steel of HMS has been based on the criteria
of strength, buckling-resistance, serviceability, and
stability requirements dictated by different codes
(e.g. AASHTO and ASD etc.). It is noteworthy
that the use of a simplified linear static model is
sufficient for the calculation of the basic response
and the eigenvalues in this case.
(2) Taiwans code uses the formula 5 . H/1000 (tip
displacement 5 . H/1000 = 15 cm) to limit building drift, its not acceptable for HMS design. The
15% deflection limitation for group-II load combination is not a serviceability requirement, but
it constitutes a safeguard for the design of HMS.
(AASHTO, 2001) In fact, reduction in deformability
is a convenient way to minimize the undesired
second-order moment effects and to ensure

continuous service of the HMS.


(3) Across-wind resonance and fatigue force are
checked as shown in Table 10. According to the
result, the critical wind velocity is equal to 2.47
m/s in the first mode, 15.45 m/s in the second
mode, and 52.05 m/s in the third mode. However,
the critical wind velocity is greater than 20 m/s;
the wind is generally too turbulent for vortex
shedding to occur. Consequently, the critical
wind speed for the first mode of vibration is less
than 50% of the top wind speed. However, one
needs to check the frequency of vortex resonance
for these higher order modes. Moreover, when
the across-wind forces are considered in HMS
design, WRD should take into account CL
(fluctuating force) instead of C L (the mean wind
force). When the towers body polygon shape has
fewer than 16 sides, the force is significant with
dC
the increasing of wind attack angle dD to the
surface of circular body. It means that the difference of wind pressure is coefficient with the
factor of attack angle.
(4) In this study, not all of the HMSs are affected by
these phenomena in WRD procedures. For
instance, windmill or wind turbine blades dont
lead to galloping.

C. W. Chien et al.: Wind-Resistant Design of High Mast Structures

VII. CONCLUSIONS
This paper has considered a number of key factors associated with the design of HMS due to the
effects of wind loading. The major findings are summarized as follows.
When increasing the number of sides on the
polygon, the total drag coefficient C d will converge
which closes to leeward drag coefficient C l , and
windward coefficient Cw will tend to 0. Based on the
simplification in CFD results, a polygon or circular
section of HMS with windward coefficient C w = 0
and an uplift coefficient CL = 0, the leeward drag coefficient C l will be close to 0.82. This result is different from Simiu, and Scanlan, (1996) For slender
towers and stacks with a circular shape in plan, it may
be assumed in all cases that Cl = 0. In this reference
C l = 0 is based on quasi-steady theory; liner fundamental modal shape.
For structural design, Cook, (1990) claimed that
windward coefficient C w = 0 is appropriate for polygons with more than eight sides, while our result suggests that polygons with more than 16 sides closely
resemble a round cylinder. And in fact, the HMS
usually takes a round shape. It seems that our result
is more proper for practical application. Therefore,
if the value of C d is adopted to 0.6 for sixteen sided
polygon, the result of along-wind force is underestimated 37% in this case study.
The design wind velocity is determined in accordance with local code and to meet the clients requirements for along-wind load of HMS. Then, it
should be compared to the total maximum response
of HMS under the critical wind velocity of vortex
resonance in along-wind and across-wind analysis for
WRD. However, if the critical wind velocity of vortex resonance is greater than 20 m/s; the wind is generally too turbulent for vortex shedding to occur.
Furthermore, the first mode is always considered in
along-wind, and the first to fourth modes should be
considered in across-wind force. As a result, dynamic
analysis using natural period and damping ratio not
only plays an important role in the detailed design of
HMS geometric configurations, it also is very important for serviceability limit states. Finally, when an
HMS is designed with the gust effect factor, G, large
than 2.33, showing flexibility by geometric configuration and structural properties, this suggests the
designer should perform a serviceability analysis required by WRD for HMS.

NOMENCLATURE
A
B
B(z)
Cd
CL
CL
Cl
Cm
Cp
Cr
CV
Cw
D

E
E MAX
Es
E(t)
F
Fd
F i(z, t)
F max
FS
Fy

F
F(z)

F(z)
f (z, t)
fs
G
g
gB
gD
gR
H
IF
IZ
Kl
Ks
M
M*
M cr
Mo
Mr
m
m1

ACKNOWLEDGMENTS
The authors would like to thank the National
Science Council of Taiwan for financial support of
this research. (Project No.: NSC 96-2211-E-019-010)

613

m(z)
N
Ng
n1

tributary projected area


maximum width of the structure or equipment
width of the structure at height z
total drag coefficient
mean of wind force
fluctuating force
leeward pressure coefficient
amplification coefficient
pressure coefficient
wind force of vortex resonance
generalized damping
windward pressure coefficient
diameter of 2/3H height
mean load effect
maximum total load effect
modulus of elasticity
load effect on the structure element
the total wind load on the structure
total drag force
randomly distributed fluctuating forces
the maximum total load effect
safety ratio factor for overturning stability
yield stress
value of the mean along-load
distributed loading at height z
mean component
fluctuating component
vortex shedding frequency
gust effect factor
peak factor for a Gaussian process
non-resonant peak factor
design peak factor
resonant peak factor
height of structure
fatigue factor
intensity of turbulence
stiffness for the first mode of the structure
mass-damping parameter = m / aD 2
primary bending moment
generalized mass
nominal moment strength in the absence
of axial load
overturning moment
resistant moment
mass of the structure
generalized mass for the first mode of the
structure
mass per unit length along the structure
number of sides of the HMS
galloping-induced vertical shear force
natural frequency for the first mode of the

614

n
P
P cr
PE
PG
P vs
p
p0
Q
q
qz
R
Re
R n, R B, R L
Sc
St
T0
Tj
Tn
t
X

X
Xa

X ave
Xc
X MAX
X max(H)
X total
x(H)
x max(H)
U
U 10
U HMS
U rel
U r(z)
U wind
u*
u wind
u(z, t)
V(z, t)
V(z cr)

V (z)
Vh

Journal of the Chinese Institute of Engineers, Vol. 33, No. 4 (2010)

structure
natural frequency
applied axial compression load
nominal strength for the axially loaded
compression member
Euler critical loading
galloping-induced vertical shear pressure
range
fatigue design of vortex shedding-induced
loads
local pressure
far upstream pressure
background response factor
velocity pressure
velocity pressure at height z
resonant response factor
Reynolds number
parameters for resonant response factor
Scruton number
Strouhal Number
period when the peak response is assumed
to occur
natural period of the structure ( j = 1, 2 ... n)
natural period of structure
thickness of shaft
random response
mean value of random response
along-wind response, e.g. maximum
along-wind displacement
mean displacement of mean along-wind
velocity at the average height point
across-wind response, e.g. maximum
across-wind displacement
maximum value of random response
total along-wind displacement
total wind response or force
average displacement caused by average
wind speed
turbulent displacement caused by the turbulence
mean value of the reference wind
basic design speed
wind velocity for the movement of each
point on the pole
relative velocity between wind and the
pole
critical wind velocity of vortex resonance
at height z
wind velocity at the nodal point
shear velocity
wind velocity at the nodal point
turbulence component
wind speed vector at any point
mean design wind speed at z cr
mean design wind speed at z
tip wind speed

vK
v oc
W
W rz
z
Zd
Z0
z cr

(z)

DL( )
DL(0)
D( )
L( )
r
s1
s2
z
a
s
x
x2
B(E)
R(E)
BF
RF
BX
RX

d
o

dC D
d

expected frequency
kinematic viscosity of air
critical wind velocity of galloping
weight of structure
across-wind force at height z
height above ground level
zero plan displacement
roughness length
at 5/6 height of structure
coefficient of the exponent in the power
law
damping ratio for vortex shedding
maximum deflection
fundamental model shape at height z
Eulers constant
X
angle = tan 1 a
Xc
coefficient of galloping force
derived value of DL( )
coefficient of along-wind force
coefficient of across-wind force
coefficient of return period
coefficient of the front side of along-wind
coefficient of the back side of along-wind
coefficient of the pressure exposure at
height z
air density
mass density of the structure
gust spectrum with standard deviation
gust spectrum with variance
R.M.S. value of the non-resonant load effect
R.M.S. value of the resonant load effect
R.M.S. value of the non-resonant load
R.M.S. value of the resonant load
R.M.S. value of the non-resonant displacement
R.M.S. value of the resonant displacement
natural circular frequency
design wind pressure
basic wind pressure
mode exponent
damping ratio of the structure
angle of wind in the main x-direction
wind attack angle
REFERENCES

AASHTO, 2001, Supports Specifications, the American Association of State Highway and Transportation Officials, Washington, USA.
ASCE-07, 2005, Minimum Design Loads for Buildings and Other Structures, American Society of
Civil Engineers, NY, USA.
China National Standard GB50009, 2001, Load Code
for the Design of Building Structures, China
Construction Industry Press, Beijing, China.

C. W. Chien et al.: Wind-Resistant Design of High Mast Structures

Chien, C. W., and Jang, J. J., 2008, Case Study of


Wind-Resistant Design and Analysis of High Mast
Structures Based on Different Wind Codes, Journal
of Marine Science and Technology, Vol. 16, No.
4, pp. 273-285
Chien, C. W., and Jang, J. J., 2008, A Study of Shape
Factors and Geometric Properties for Wind-Resistant Design of High Mast Structures, The 2nd
National Conference on Wind Engineering, Taipei,
Taiwan, R.O.C.
Chien, C. W., and Jang, J. J., 2005, Wind-Resistant
Design of High Mast Structures, Asia Pacific
of Engineering Science and Technology, Vol. 3,
No. 2, pp. 549-570. (in Chinese)
Cook, N. J., 1990, The Designers Guide to Wind
Loading of Buildings. Part 2: Static Structures,
Butterworths Press, London, England.
Davenport, A. G., 1964, Note on the Distribution of
the Largest Value of a Random Function with
Application to Gust Loading, Proceed of the Institute of Civil Engineers, London, Vol. 28, Issue
2, pp. 187-196.
Holmes, J. D., 2001, Wind Loading of Structures, 3rd
ed., Spon Press, London, UK.
Germano, M., Piomelli, U., Moin, P., and Cabot, W.
H., 1991, A Dynamic Subgrid-Scale Eddy Viscosity Model, Physics of Fluids, Vol. 3, Issue 7,
pp. 1760-1765.
John, W. van de Lindt, and Jonathan, S. Goode, 2006,
Development of a Reliability-based Design Procedure for High-Mast Lighting Structural Supports
in Colorado, Colorado Department of Transportation Safety and Traffic Engineering Branch and
Staff Bridge Branch, No. SRR-91, CO, USA.
Kareem, A., 1981, Wind-Excited Response of Buildings in Higher Modes, Journal of Structural Engineering ASCE, Vol. 107, No. ST4, pp. 701-706
Kevin, W., Johns, etc., 1998, The Development of
Fatigue Design Load Ranges for Cantilevered Sign
and Signal Support Structures, Journal of the Wind
Engineering and Industrial Aerodynamics, Vol.
71-78, No. 77-78, pp. 315-326.
Melbourne, W. H., 1997, Predicting the Across-Wind
Response of Masts and Structure Members,
Journal of the Wind Engineering and Industrial
Aerodynamics, Vol. 69-71, No. 69-71, pp. 91-103.
Mendis, P., and Cheung, J., 2007, Wind Loading on

615

Tall Building, EJSE Special Issue: Loading on


Structures, pp. 41-54.
SAP2000, Computers and Structures Inc., Berkeley,
CA, USA.
Schilling, C. G., 1965, Buckling Strength of Circular Tubes, Journal of the Structural Division, Vol.
19, No. 5. American Society of Civil Engineers,
NY, USA, pp. 325-348.
Scruton, C., 1963, On the Wind Excited Oscillations
of Stacks, Towers and Masts, Proceedings Wind
Effects on Buildings and Structures, HM Stationary
Office, London, UK, pp. 798-836.
Simiu, E., and Scanlan, R. H., 1996, Wind Effects on
Structures, 3rd ed., John Wiley and Sons, NY,
USA.
Smith, J. W., 1988, Vibration of Structures Application in Civil Engineering Design, Chapman and
Hall, NY, USA.
Solaria, G., 1982, Alongwind Response Estimation:
Closed Form Solution, ASCE Journal of the Structural Division, Vol. 108, No. 1, pp. 225-244.
Solaria, G., 1999, Gust Buffeting and Aeroelastic
Behavior of Poles and Monotubular Tower,
Journal of the Fluids and Structures, No. 13, pp.
877-905.
Song, C. C. S., and Yuan, M., 1998, A Weakly Compressible Flow Model and Rapid Convergence
Methods, Journal of Fluids Engineering, Vol.
110, No. 4, pp. 441-455.
Tsai, I -C., Cherng, R. -H., and Hsiang, W. B., 2006,
Establishment of Building Wind loading Code and
the associated Commentary, Building Research
Institute of Ministry of Internal Affairs, Taipei,
R.O.C.
Wilson, E. L., and Habibullah, A., 1987, Static and
Dynamic Analysis of Multi-Story Buildings Including P-delta Effects, Earthquake Spectra,
Earthquake Engineering Research Institute, Vol.
3, No. 3, pp. 289-98.
Wang, Z. M., 1995, Tower structure design manual,
China Construction Industry Press, Shanghai,
China.
Manuscript Received: Jul. 04, 2008
Revision Received: Apr. 21, 2009
and Accepted: May 21, 2009

You might also like