Вы находитесь на странице: 1из 19

Journal of Wind Engineering

and Industrial Aerodynamics 95 (2007) 119


Investigation of closed-form solutions to estimate
fatigue damage on a building
Nag-Ho Ko
a,
, Young-Moon Kim
b
a
Hanseok Engineering Consultants Inc., Seoul 138-843, Korea
b
Chonbuk National University, Chonju 561-756, Korea
Received 2 September 2005; received in revised form 28 February 2006; accepted 4 April 2006
Available online 12 May 2006
Abstract
If a stress process is wide-band, it is not obvious what constitutes a cycle and how cycles should be
counted so that the Miners rule can be employed. If a stress process is non-Gaussian, the stress
process may cause accelerated fatigue damage. In this study, high-cycle fatigue damage for
fasteners of curtain walls on a side face of a square building is estimated by not only closed-form
solutions that consider the effects of wide-band and non-Gaussian characteristics, but also rainow
analysis that uses rainow cycle counting method to determine the number of cycles and the
magnitude of stress amplitude or range that are needed to employ Miners rule. The fatigue damage
estimations are compared to investigate the applicability of the closed-form solutions for initial
design calculations.
r 2006 Elsevier Ltd. All rights reserved.
Keywords: Local wind pressure; Square; Building; Fatigue damage; Wide-band; Non-gaussian; Closed-form
solution; Rainow analysis
1. Introduction
Components of machines, vehicles, and structures are frequently subjected to repeated
loads, also called cyclic loads, and the resulting cyclic stresses can lead to micro-
scopic physical damage to the materials involved. Even at stresses well below a given
materials ultimate strength, this microscopic damage can accumulate with continued
ARTICLE IN PRESS
www.elsevier.com/locate/jweia
0167-6105/$ - see front matter r 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jweia.2006.04.001

Corresponding author. Tel.: +82 2 2185 8375; fax: +82 2 417 3908.
E-mail address: knagho@yahoo.co.kr (N.-H. Ko).
cycling until it develops into a crack or other macroscopic damage that leads to failure
of the component. This process of damage and failure due to cyclic loading is called
fatigue [1].
The uctuating nature of wind loading can produce uctuating stresses in struc-
tures. When the stress process is stationary narrow-band Gaussian and the stress
amplitudes have a Rayleigh Distribution, fatigue damage can be estimated from the widely
used form that was rst proposed by Miles [2,3]. However, when the stress process is
stationary wide-band non-Gaussian, fatigue damage cannot be immediately estimated.
Then the effects of wide-band and non-Gaussian characteristics on fatigue damage have to
be considered.
Davenport [4] proposed a method to estimate wind-induced fatigue damage by
assuming the wind loading as a wide-band Gaussian process. Patel and Freathy [5] pro-
posed a simplied method for assessing wind-induced fatigue damage under the
assumption of linear relationship between the stress and wind uctuations. Wyatt [6]
assessed the sensitivity of lattice towers to fatigue induced by wind gusts. Lynn and
Stathopoulos [7] improved the predictions of fatigue damage by including the effect
of non-Gaussian distribution by using a mixed GaussianWeibull extreme model.
Winterstein [8,9] derived Hermite moment models to predict the ratio of non-Gaussian
fatigue damage estimates to Gaussian fatigue damage estimates. Gerhardt and
Kramer [10] proposed load cycle distributions for proof testing of building components
against wind-induced fatigue for temperate climatic regions. Xu [1113] investigated
wind-induced fatigue damage of roof cladding. Letchford and Norville [14] developed a
wind pressure loading cycles for wall cladding during hurricanes that was seen to
lead to fatigue damage. Jancauskas et al. [15] incorporated analytical models into
computer programs to simulate fatigue behavior of roof cladding during the passage of a
tropical cyclone. Mahendran [16] proposed a random block loading based on fatigue wind
loading matrices obtained from wind tunnel testing and computer modeling for the
simulation of cyclonic wind forces on roof cladding. Mahendran [17] proposed a new
fatigue loading sequence for codication after considering the wind speed as well as wind
direction variation during a tropical cyclone. Mikitarenko and Perelmuter [18] proposed a
method to estimate the safe fatigue life of steel towers under the action of wind
uctuations. Holmes [19] derived closed-form expressions for upper and lower limits on
fatigue life under along-wind loading. Repetto and Solari [2022] investigated alongwind,
crosswind, and directional wind-induced fatigue damage of slender vertical structures.
Wyatt [23] determined gust action stress cycle counts for fatigue checking of line-like
steel structures.
In this study, high-cycle fatigue damage for fasteners that x curtain walls on a side
face of a square building is estimated by closed-form solutions and rainow analysis,
and the fatigue damage estimations are compared to investigate the applicability of
the closed-form solutions for initial design calculations. The closed-form solutions
consider the effects of wide-band characteristics and non-Gaussian probability distribu-
tions on the fatigue damage for an equivalent narrow-band Gaussian stress process.
The closed-form solutions consist of four closed-form expressions using two correction
factors for wide-band effects and two correction factors for non-Gaussian effects.
In rainow analysis, the rainow cycle counting method and Miners rule are applied
to a stress time histories with wide-band non-Gaussian distribution to estimate fatigue
damage.
ARTICLE IN PRESS
N.-H. Ko, Y.-M. Kim / J. Wind Eng. Ind. Aerodyn. 95 (2007) 119 2
2. Fatigue damage under variable amplitude loading
If a test specimen of a material or an engineering component is subjected to a sufciently
severe cyclic stress, a fatigue crack or other damage will develop, leading to complete
failure of the member. The results of such tests from a number of different stress levels may
be plotted to obtain a stresslife curve, also called SN curve. There is a signicant benet
in simplicity of analysis if it can be assumed that fatigue strength can be expressed by a
linear function in loglog space. If so, the empirical form is implied as follows:
NS
m
A, (1)
where N is the number of cycles until failure, S the stress amplitude or range that depends
on the choice of the analyst, A the fatigue strength coefcient that depends on the material,
and m the fatigue strength exponent that depends on the material. It should be noted that
either S or A can be based on amplitude or range, but the distinction is important.
A linear damage rule that was originally suggested by Palmgren and was developed by
Miner [24] has been used to estimate fatigue damage because of its simplicity and
usefulness [25]. Miners rule simply states that fatigue failure is expected when the total
fatigue damage, D in Eq. (2), sums to unity
D

k
i1
n
i
N
i
, (2)
where n
i
is the number of applied cycles at a stress amplitude S
i
, and N
i
the number of
cycles to failure from the SN curve at a stress amplitude S
i
.
2.1. Fatigue damage under narrow-band Gaussian stress process
When a stress process is narrow-band Gaussian distribution and the stress amplitudes
are Rayleigh distribution, fatigue damage can be estimated from the equations that were
rst proposed by Miles [2]. When A and S are based on amplitude, fatigue damage at time
t is as follows:
D
n

0
t
A

2
p
s
s

m
G
m
2
1
_ _
. (3)
When A and S are based on range, fatigue damage at time t is changed as follows:
D
n

0
t
A
2

2
p
s
s

m
G
m
2
1
_ _
, (4)
where n

0
is the rate of crossing of the mean stress with positive slope, s
s
the standard
deviation of the entire stress process, and m the fatigue strength exponent.
2.2. Fatigue damage for an equivalent narrow-band Gaussian stress process
Stress cycles are easily identied for the narrow-band process having Rayleigh peaks.
But if a stress process is wide-band process, it is not immediately obvious how to count
stress cycles to be used with Miners rule. One approximation is to assume that fatigue
damage for a wide-band Gaussian stress process will be the same fatigue damage as the
narrow-band Gaussian process with the same rate of crossing of the mean stress with
ARTICLE IN PRESS
N.-H. Ko, Y.-M. Kim / J. Wind Eng. Ind. Aerodyn. 95 (2007) 119 3
positive slope and the same standard deviation. Fortunately, this algorithm leads to a
conservative damage estimate [26]. Thus, fatigue damage at time t under a wide-band
Gaussian stress process can be estimated using an equivalent narrow-band Gaussian stress
process as follows:
D
NB

n

0
t
A

2
p
s
s
_ _
m
G
1
2
m 1
_ _
: A and S are based on amplitude;
D
NB

n

0
t
A
2

2
p
s
s
_ _
m
G
1
2
m 1
_ _
: A and S are based on range;
(5)
where D
NB
is the fatigue damage for an equivalent narrow-band Gaussian stress process.
2.3. Rainow analysis
Dowling [27] and Nelson [28] have provided summaries of cycle counting methods, i.e.,
schemes for obtaining statistics of stress ranges for fatigue prediction when the stress
histories are wide-band. Dowling [27] stated that the rainow cycle counting method is
generally regarded as the method leading to better predictions of fatigue life. It can identify
events that are compatible with constant range fatigue data in a variable stress sequence.
The rainow cycle counting method determines the number of cycles and the magnitude of
stress range from variable stress histories. At present a number of rainow cycle counting
techniques are in use. In this study, the standardization of rainow cycle counting method
by Amzallag et al. [29] is adopted to count the number of cycles and the magnitude of
stress range present in variable stress time histories.
In rainow analysis, rainow cycle counting method and Miners rule are used to estimate
fatigue damage. The stress amplitude or range depends on the choice of the analyst. Rainow
cycle counting method is used to determine the number of cycles and the magnitude of stress
amplitude or range in a wide-band stress process. The number of applied cycles at a stress
amplitude, or range, S
i
, and N
i
, the number of cycles to failure at a stress amplitude S
i
from
the SN curve, are used to estimate fatigue damage by employing Miners rule, Eq. (2).
However, the rainow analysis requires extensive numerical calculations.
3. Closed-form solutions on a side face of a square building
3.1. Wide-band effects
The Rayleigh approximation is simple and widely known, and involves assumptions that
are reasonable for a narrow-band Gaussian process. Thus, it may be helpful to characterize
any other method for analyzing a wide-band process in terms of how much its fatigue
damage differs from that of the Rayleigh approximation.
Wirsching and Light [30] proposed fatigue damage, D, under a wide-band stresses as follows:
D lb; mD
NB
, (6)
where D
NB
is fatigue damage for an equivalent narrow-band Gaussian stress process,
Eq. (5) and lb; m is the correction factor as follows:
lb; m a 1 a1 b
b
, (7)
ARTICLE IN PRESS
N.-H. Ko, Y.-M. Kim / J. Wind Eng. Ind. Aerodyn. 95 (2007) 119 4
where a and b are functions of the fatigue strength exponent m, in Eq. (1)
a 0:926 0:033m, (8)
b 1:587m 2:323, (9)
where b is a spectral bandwidth parameter.
Ortiz and Chen [31] and Perng [32] proposed fatigue damage, D, under a wide-band
stresses as follows:
D l
k
D
NB
, (10)
where D
NB
is fatigue damage for an equivalent narrow-band Gaussian stress process,
Eq. (5) and l
k
the correction factor as follows:
l
k

b
m
k
e
, (11)
where m is the fatigue strength exponent in Eq. (1), b
k
the generalized spectral bandwidth,
and e the irregularity factor as follows:
b
k

M
2
M
k
M
0
M
k2

, (12)
k
2:0
m
, (13)
e
M
2

M
0
M
4
p (14)
and where, M
j
is the jth moment of the one-sided spectral density function for stress
process, W(f) as follows:
M
j

_
1
0
f
j
Wf df : (15)
3.2. Non-Gaussian effects
If the stress process has a greater than Gaussian probability of taking on large values,
then this is likely to cause large stress ranges, and these may cause accelerated fatigue
damage [26,33]. In order to illustrate the non-Gaussian effects on fatigue, the ratio of non-
Gaussian fatigue damage estimates to Gaussian fatigue damage estimates was introduced
as follows:
g
ED
NG

ED
G

, (16)
where D
NG
is fatigue damage for non-Gaussian distribution and D
G
is fatigue damage for
Gaussian distribution.
Winterstein [8] offered a simplied g using the rst-order softening Hermite model in
1985 as follows:
g 1
mm 1kurtosis 3
24
_ _
for kurtosis43; (17)
ARTICLE IN PRESS
N.-H. Ko, Y.-M. Kim / J. Wind Eng. Ind. Aerodyn. 95 (2007) 119 5
g 1
mm 1kurtosis 3
24
_ _
1
for kurtosiso3, (18)
where m is the fatigue strength exponent in Eq. (1).
Winterstein [9] offered another simplied g using the second-order softening Hermite
model in 1988 as follows:
g

p
p
a
2V
R
!
_ _
m
mV
R
!
m=2
_ _
!

p
p
a
2GV
R
1
_ _
m
GmV
R
1
G m=2 1
_ _
;
(19)
where
V
R

41 h

4
h
4
p
p
_
;
a
1

1 2h
2
3
6h
2
4
_ ;
h

4

kurtosis 3
24
;
h
3

skewness
4 2

1 1:5kurtosis 3
_ ;
h
4

1 1:5kurtosis 3
_
1
18
and where G is Gamma function and m is the fatigue strength exponent in Eq. (1).
3.3. Fatigue damage under wide-band non-Gaussian stress process
From investigations to the offshore structures subjected to wave loadings, if the stress is
neither Gaussian nor narrow-band, it is sometimes acceptable to use correction factors for
the wide-band effects and non-Gaussian effects [26].
In this study, because the stresses on fasteners of curtain wall on a side face of a square
building are assumed to be wide-band non-Gaussian stresses process, Eq. (20) is pro-
posed as a closed-form solution for estimating fatigue damage. The correction factors
for wide-band effects used in this paper are based on the effect of narrow-band versus
wide-band processes. The correction factors for non-Gaussian effects used in this paper
are based on the effect of Gaussian versus non-Gaussian processes. Therefore, it is
assumed that the wide-band and non-Gaussian effects are independent in the closed-form
solutions.
D lgD
NB
, (20)
where l is the correction factor for considering wide-band effects using Eq. (7) or Eq. (11),
g the correction factor for considering non-Gaussian effects using Eq. (17) or Eq. (19), and
D
NB
is fatigue damage for an equivalent narrow-band Gaussian stress process.
ARTICLE IN PRESS
N.-H. Ko, Y.-M. Kim / J. Wind Eng. Ind. Aerodyn. 95 (2007) 119 6
3.4. Closed-form expressions on a side face of a square building
In this study, four closed-form expressions are proposed to estimate wind-induced
fatigue damage of fasteners on a side face of a square building.
The rst closed-form expression is that the correction factor for considering wide-band
effects, l, uses Eq. (7) and the correction factor for considering non-Gaussian effects, g,
uses Eq. (17) in Eq. (20).
The second closed-form expression is that the correction factor for considering wide-
band effects, l, uses Eq. (7) and the correction factor for considering non-Gaussian effects,
g, uses Eq. (19) in Eq. (20).
The third closed-form expression is that the correction factor for considering wide-band
effects, l, uses Eq. (11) and the correction factor for considering non-Gaussian effects, g,
uses Eq. (17) in Eq. (20).
The fourth closed-form expression is that the correction factor for considering wide-
band effects, l, uses Eq. (11) and the correction factor for considering non-Gaussian
effects, g, uses Eq. (19) in Eq. (20).
4. Results
4.1. Wind tunnel test
The experiments were carried out in the boundary layer wind tunnel at Chonbuk
National University, Chonju, Korea [34]. The working section of the tunnel is 14 m in
length, 1.5 m in width and 1.2 m in height. The reference velocity was taken at the top of
the model and was approximately 6 m/s in wind tunnel. The velocity scale for this study
was set out to be approximately 1:5. The power spectral density for the longitudinal
component of turbulence was adjusted to that of von Karman. Plexiglass model of a
square building was tested in urban terrain condition (a 0.30) for zero wind angles.
The dimensions of the model used in this study are reported in Table 1. In this study, 23
taps in Fig. 1, which are located on a side face, were used to measure local wind pressure
uctuations. The taps located near to the each edge are at a distance of 1 cm from the each
edge of the model. The taps located on center are at a distance of 5 cm from the edges of
the model. S8 and S10 are located at a distance of 3 cm from their near edge of the model.
The height of the taps is represented in Table 2.
Pressure data were measured for duration of 44.992 s at a sampling rate of 250 samples/s
at all taps in Fig. 1. At a length ratio of 1:400 and mean velocity ratio of 1:5, the resulting
time history of 44.992 s providing statistically stable mean and variance is equivalent to
about 1 h in full scale. The measured pressure signals were converted to pressure coefcient
signals by dividing them by the reference dynamic pressure at the top of the model.
ARTICLE IN PRESS
Table 1
Model used in this study
Breadth (cm) Depth (cm) Height (cm) Side ratio Aspect ratio Blockage ratio (%)
10 10 40 1:1 4:1 2.2
N.-H. Ko, Y.-M. Kim / J. Wind Eng. Ind. Aerodyn. 95 (2007) 119 7
Table 2 shows the statistics, mean, variance, skewness, and kurtosis, of the measured
pressure coefcient signals at all the taps in Fig. 1. The measured pressure coefcient
signals on a side face of a square building have negative means. The mean values of the
measured pressure coefcient signals are different at each tap. The intensity of uctuations
(variance) of the measured pressure coefcient signals is different at each tap. According to
the skewness and kurtosis values shown in Table 2, the local wind pressure uctuations on
a side face of a square building follow either moderately non-Gaussian or highly non-
Gaussian distribution.
It should be noted that high uctuations (variance) do not guarantee highly non-
Gaussian (high skewness and kurtosis values) in a signal. For instance, though sample S2 is
highly uctuating, the uctuations are approximately evenly distributed above and below
ARTICLE IN PRESS
Fig. 1. Location of pressure taps and wind direction.
N.-H. Ko, Y.-M. Kim / J. Wind Eng. Ind. Aerodyn. 95 (2007) 119 8
the mean, and it is not highly non-Gaussian. Sample S23 is highly non-Gaussian, but it has
low value of variance compared to sample S2. This shows that there is no clear evidence
concerning any relation between the intensity of uctuations (variance) and the non-
Gaussian distribution of a signal.
The measured signals at S5, S11, S13, and S18 are presented in the form of pressure
coefcient (C
p
) versus time in Fig. 2. They seem to have highly negatively going spikes that
are responsible for its non-Gaussian nature. It seems that two properties of spikes, namely
the frequency of occurrence and the magnitude, affect the non-Gaussian nature.
In Fig. 3, showing the PDF (probability density function), the abscissa represents
normalized pressure coefcient in the form of (C
p
C
p
,mean)/C
p
,rms (C
p
,mean
corresponds to mean pressure coefcient, and C
p
,rms corresponds to root mean square
pressure coefcient). The probability ordinates are shown on a logarithm scale in order to
show the tail end of the PDF clearly. The PDF gures indicate much higher probability for
the larger negative pressure coefcients than a Gaussian distribution would predict. The
distributions of the measured pressure coefcient signals have the mode of the distribution
near zero that is shifted slightly to the positive side, and is located at a value greater than
the mean. Therefore, the skewnesses of the measured pressure coefcient signals are less
than zero [35].
ARTICLE IN PRESS
Table 2
Statistics of the measured pressure coefcient signals
Location (about) Tap No. Mean Variance Skewness Kurtosis
0.95 height of the model S1 0.98 0.11 0.66 3.55
S2 0.84 0.12 0.33 3.27
S3 0.48 0.07 0.67 4.08
0.9 height of the model S4 0.88 0.09 0.55 3.34
S5 0.89 0.13 0.49 3.44
S6 0.57 0.09 0.44 3.26
0.8 height of the model S7 0.83 0.09 0.62 3.41
S8 0.87 0.12 0.70 3.91
S9 0.87 0.13 0.61 3.63
S10 0.76 0.13 0.63 3.65
S11 0.67 0.13 0.60 3.71
0.6 height of the model S12 0.84 0.11 0.78 3.96
S13 0.84 0.15 0.76 3.92
S14 0.62 0.13 0.70 3.63
0.4 height of the model S15 0.85 0.12 0.96 4.42
S16 0.77 0.13 0.66 3.68
S17 0.54 0.11 1.03 4.70
0.2 height of the model S18 0.83 0.15 1.11 4.73
S19 0.64 0.13 1.05 5.14
S20 0.45 0.08 1.07 4.57
0.1 height of the model S21 0.73 0.11 1.04 4.93
S22 0.49 0.10 1.16 5.47
S23 0.31 0.03 0.10 4.98
N.-H. Ko, Y.-M. Kim / J. Wind Eng. Ind. Aerodyn. 95 (2007) 119 9
4.2. Non-Gaussian simulation based on spectral correction approach
Considering the stochastic properties of local wind pressure uctuations, a number of
time histories of the local wind pressure uctuations are needed to carry out fatigue
damage estimation. Wind tunnel test has been used to measure local wind pressure
uctuations but it is expensive and time consuming in collecting a number of time histories
of local wind pressure uctuations. Therefore, it may be useful to utilize a suitable
simulation method for generation of local wind pressure uctuations. Then non-Gaussian
local wind pressure signals are simulated by spectral correction (SC) approach [36,37] in
this study. The number of the simulated local wind pressure signals that correspond to
their measured local wind pressure signal is 100 samples at each tap in Fig. 1.
There are two options for the exit criterion in the iterative SC algorithm. In the rst exit
option, the loop stops after the spectral correction section if the error between the target
moments and resulting moments is within the specied tolerance. In the second exit option,
the loop stops after the moment correction section if the error between the target spectrum
and resulting spectrum is within the specied tolerance.
ARTICLE IN PRESS
(a)
(b)
(d) (c)
Fig. 2. Measured pressure coefcient signals. (a) Measured pressure coefcient signal at S5. (b) Measured
pressure coefcient signal at S11. (c) Measured pressure coefcient signal at S13. (d) Measured pressure coefcient
signal at S18.
N.-H. Ko, Y.-M. Kim / J. Wind Eng. Ind. Aerodyn. 95 (2007) 119 10
The rst exit option is used in this study. Therefore, a check is performed on the error
between the target skewness and kurtosis, g
3
T
,g
4
T
, and the resulting simulation skewness and
kurtosis, g
3
, g
4
as Eq. (21). If the error is within the specied tolerance, the algorithm stops
and the non-Gaussian process with target spectrum G
T
is produced.
error
g
T
3
g
3
g
T
3

g
T
4
g
4
g
T
4

. (21)
The specied tolerance used in this study is 1.0 10
5
. Therefore, the mean and variance
of the simulated local wind pressure coefcient signals coincide with those of their
corresponding measured local wind pressure coefcient signal at each tap in Fig. 1. The
skewness and kurtosis of the simulated local wind pressure coefcient signals almost
coincide with those of their corresponding measured local wind pressure coefcient signal
within four places of decimals at the taps. The power spectrum of the simulated local wind
pressure coefcient signals coincides with their corresponding target spectrum, namely, the
ARTICLE IN PRESS
-7 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5
10
-1
10
-2
10
-3
(Cp-Cp,mean)/Cp,rms (Cp-Cp,mean)/Cp,rms
P
r
o
b
a
b
i
l
i
t
y
10
-1
10
-2
10
-3
P
r
o
b
a
b
i
l
i
t
y
-7
10
-1
10
-2
10
-3
P
r
o
b
a
b
i
l
i
t
y
10
-1
10
-2
10
-3
P
r
o
b
a
b
i
l
i
t
y
Measured S5
Simulated S5
Normalized Gaussian
-6 -5 -4 -3 -2 -1 0 1 2 3 4 5
-7 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5
(Cp-Cp,mean)/Cp,rms (Cp-Cp,mean)/Cp,rms
-7 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5
Measured S11
Simulated S11
Normalized Gaussian
Measured S13
Simulated S13
Normalized Gaussian
Measured S18
Simulated S18
Normalized Gaussian
(a) (b)
(c) (d)
Fig. 3. PDF plots of the measured and corresponding simulated pressure coefcient signals. (a) PDF of the
measured and a corresponding simulated pressure coefcient signal at S5. (b) PDF of the measured and a
corresponding simulated pressure coefcient signal at S11. (c) PDF of the measured and a corresponding
simulated pressure coefcient signal at S13. (d) PDF of the measured and a corresponding simulated pressure
coefcient signal at S18.
N.-H. Ko, Y.-M. Kim / J. Wind Eng. Ind. Aerodyn. 95 (2007) 119 11
spectrum of their corresponding measured local wind pressure coefcient signal at each tap
in Fig. 1.
One signal of 100 simulated local wind pressure coefcient signals at S5, S11, S13, and
S18 is presented in the form of pressure coefcient (C
p
) versus time in Fig. 4, respectively.
The simulated signals appear similar to the corresponding measured signal at all the taps as
the representative examples shown in Fig. 4.
The PDF plot for one of 100 simulated signals at S5, S11, S13, and S18 is presented
along with for their corresponding measured signals in Fig. 3, respectively. The PDF plots
between the measured and corresponding simulated signals match well except negligible
discrepancies at the negative tail end at all the taps as the representative examples shown
in Fig. 3.
4.3. Stress analysis in a fastener of curtain wall
Because the dimensions of cladding panels are relatively small (a few meters at most), it
is acceptable to assume that the uctuating loads acting over the area of the panel are
uniformly distributed at any one instant and proportional to the pressures at the panel
ARTICLE IN PRESS
(a)
(b)
(d) (c)
Fig. 4. Simulated pressure coefcient signals. (a) A simulated pressure coefcient signal at S5. (b) A simulated
pressure coefcient signal at S11. (c) A simulated pressure coefcient signal at S13. (d) A simulated pressure
coefcient signal at S18.
N.-H. Ko, Y.-M. Kim / J. Wind Eng. Ind. Aerodyn. 95 (2007) 119 12
center [38]. The action of wind on parts and portions of the exterior envelope on which
loads are to be determined, such as components of curtain walls and roof cladding systems,
can in many situations be treated as quasistatic. A quasistatic load implies that the internal
stresses in the portion under investigation are in direct proportion to the applied force and
that the part is not in signicant resonant motion at one or more of its natural frequencies.
The reason for the quasistatic response is that the frequencies of vibration of such
components are usually well above those at which there is appreciable energy in the
turbulent ow eld. Portions with natural frequencies above approximately 12 Hz are
likely to respond quasistatically [39]. If the natural frequency of the cladding panel is high,
the dynamic effects of wind on the cladding panel can be disregarded. The cladding glass
used in buildings varies in natural frequency from about 10 to 50 Hz, and they have
suffered no damage from dynamic effects of wind [40]. Allen and Dalgliesh [41] proposed
that a cladding panel could be represented as a statically loaded structure and that
dynamic amplication could be neglected.
In this study, it is assumed that wind loading on a fastener, a component of curtain wall,
has uctuating character similar to the uctuating loads acting over the area of a cladding
glass. Therefore, the maximum stress S(t) in a fastener due to bending is given by
St CQt
f
, (22)
where f is equal to unity because stress is assumed to be directly proportional to load.
Therefore, C is in terms of the basic beam stress formula as follows:
C
L
I
c, (23)
where L is the moment arm to Q(t), I the area moment of inertia about the bending axis,
and c the distance from neutral axis to edge. In Eq. (22), Q(t) is the wind load on a fastener
with a tributary area (A
f
) that xes curtain walls located on a side face of a square building.
Qt
1
2
rU
2
C
p
tA
f
, (24)
where r is the air density (taken to be 1.22625 kg/m
3
in this study), U the mean wind
velocity at the reference height (taken to be 30 m/s in this study), C
P
(t) the wind pressure
coefcients at time t, and A
f
the tributary area of a fastener.
4.4. Example
It can be thought that the stress processes of the critical section for high-cycle fatigue
failures of a fastener subjected to uctuating wind pressures are equivalent to those of
the xed end of a cantilever beam. Therefore, a cantilever beam is regarded as a
component like a fastener that xes curtain wall located on a side face of a square building
in this example.
Consider a cantilever beam with a length L 8 cm, a width b 10 cm, a thickness
h 1 cm, and a concentrated load Q(t) applied at the end. Q(t) is obtained by Eq. (24) for
which C
p
(t) are the simulated local wind pressure coefcient time histories at each tap in
Fig. 1, r 1.22625 kg/m
3
, U 30 m/s, and A
f
6.3 m
2
. In Eq. (23), C 6L/bh
2
.
The material used in ASTM A-36; mean yield strength 290 MPa, mean ultimate
strength 470 MPa. The straight-line portion of the SN curve based on range is
ARTICLE IN PRESS
N.-H. Ko, Y.-M. Kim / J. Wind Eng. Ind. Aerodyn. 95 (2007) 119 13
modeled as a linear model, NS
m
A, for which m 4.895, and A 1.84 10
17
MPa
m
[42].
4.5. Results
High-cycle fatigue damage is estimated by the closed-form solutions and rainow
analysis, and the fatigue damage estimations are compared to investigate the applicability
of the closed-form solutions at each tap in Fig. 1.
In this study, the stress process has a nonzero mean at each tap but the inuence of the
mean on fatigue damage is not considered.
In the closed-form solutions, the correction factors for wide-band, l (b,m), Eq. (7),
proposed by Wirsching and Light [30] or l
k
, Eq. (11), proposed by Ortiz and Chen [31] and
Perng [32], are used to consider the wide-band effects on fatigue damage for an equivalent
narrow-band Gaussian stress process. The correction factors for non-Gaussian distribu-
tions, g, Eq. (17), proposed by Winterstein [8] or g, Eq. (19), proposed by Winterstein [9]
are used to consider the non-Gaussian effects on fatigue damage for an equivalent narrow-
band Gaussian stress process.
Because the 100 simulated local wind pressure signals at a tap have the same rate of
crossing of the mean stress with positive slope and the same standard deviation as one
another and the stresses on fasteners are assumed to be directly proportional to wind load,
the fatigue damage which the closed-form solution estimates is the same value in the 100
simulated local wind pressure signals at the tap.
Fig. 5 shows the fatigue damage that the rst closed-form expression estimates, and the
fatigue damage which rainow analysis estimates for the 100 simulated local wind pressure
signals at each tap in Fig. 1.
Fig. 6 shows the fatigue damage that the second closed-form expression estimates, and
the fatigue damage which rainow analysis estimates for the 100 simulated local wind
pressure signals at each tap in Fig. 1.
Fig. 7 shows the fatigue damage that the third closed-form expression estimates, and the
fatigue damage which rainow analysis estimates for the 100 simulated local wind pressure
signals at each tap in Fig. 1.
Fig. 8 shows the fatigue damage that the fourth closed-form expression estimates, and
the fatigue damage which rainow analysis estimates for the 100 simulated local wind
pressure signals at each tap in Fig. 1.
The closed-form expressions with the correction factor for wide-band effects proposed
by Wirsching and Light and the correction factor for non-Gaussian effects using the rst-
order softening Hermite model and the second-order softening Hermite model, the rst
closed-form expression and the second closed-form expression, estimate larger fatigue
damage than the upper limits of the fatigue damage which the rainow analysis estimates
at all the taps as shown in Figs. 5 and 6.
The closed-form expression with the correction factor for wide-band effects proposed by
Ortiz and Chen and Perng and the correction factor for non-Gaussian effects using the
rst-order softening Hermite model, the third closed-form expression, estimates larger
fatigue damage than the upper limits of the fatigue damage which rainow analysis
estimates at all the taps except S2 as shown in Fig. 7 and Table 3. The difference at S2 is
very small.
ARTICLE IN PRESS
N.-H. Ko, Y.-M. Kim / J. Wind Eng. Ind. Aerodyn. 95 (2007) 119 14
ARTICLE IN PRESS
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
Rainflow analysis + The second closed-form expression
Tap Locations
F
a
t
i
g
u
e

D
a
m
a
g
e

x

0
.
0
0
0
1
D 2nd cfe
Upper of Drf
Average of Drf
Lower of Drf
Fig. 6. Comparison of the fatigue damage estimated using rainow analysis with the fatigue damage estimated
using the second closed-form expression to the 100 simulated signals.
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
Rainflow analysis + The first closed-form expression
Tap Locations
F
a
t
i
g
u
e

D
a
m
a
g
e

x

0
.
0
0
0
1
D 1st cfe
Upper of Drf
Average of Drf
Lower of Drf
Fig. 5. Comparison of the fatigue damage estimated using rainow analysis with the fatigue damage estimated
using the rst closed-form expression to the 100 simulated signals.
N.-H. Ko, Y.-M. Kim / J. Wind Eng. Ind. Aerodyn. 95 (2007) 119 15
ARTICLE IN PRESS
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
Rainflow analysis + The third closed-form expression
Tap Locations
F
a
t
i
g
u
e

D
a
m
a
g
e

x

0
.
0
0
0
1
D 3rd cfe
Upper of Drf
Average of Drf
Lower of Drf
Fig. 7. Comparison of the fatigue damage estimated using rainow analysis with the fatigue damage estimated
using the third closed-form expression to the 100 simulated signals.
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
Rainflow analysis + The fourth closed-form expression
Tap Locations
F
a
t
i
g
u
e

D
a
m
a
g
e

x

0
.
0
0
0
1
D 4th cfe
Upper of Drf
Average of Drf
Lower of Drf
Fig. 8. Comparison of the fatigue damage estimated using rainow analysis with the fatigue damage estimated
using the fourth closed-form expression to the 100 simulated signals.
N.-H. Ko, Y.-M. Kim / J. Wind Eng. Ind. Aerodyn. 95 (2007) 119 16
The closed-form expression with the correction factor for wide-band effects proposed by
Ortiz and Chen and Perng and the correction factor for non-Gaussian effects using the
second-order softening Hermite model, the fourth closed-form expression, estimates larger
fatigue damage than the upper limits of the fatigue damage which rainow analysis
estimates at all the taps except S19, S21, S22, and S23 as shown in Fig. 8 and Table 3. The
differences at S19, S21, S22, S23 are very small.
From Figs. 58, and Table 3, it is shown that the correction factor for wide-band effects
proposed by Ortiz and Chen and Perng estimates better fatigue damage than the correction
factor for wide-band effects proposed by Wirsching and Light, and the second order
Hermite model estimates better fatigue damage for highly non-Gaussian wind pressure
time series than the rstorder Hermite model in the correction factor for non-Gaussian
effects.
5. Conclusions
The fatigue damage has been estimated by the closed-form solutions and rainow
analysis. The fatigue damage estimations have been compared to investigate the
applicability of the closed-form solutions that estimate wind-induced fatigue damage on
a side face of a square building.
The closed-form solutions estimate larger fatigue damage than the average fatigue
damage which rainow analysis estimates at all the taps in Fig. 1. The rst closed-form
expression and the second closed-form expression estimate larger fatigue damage than the
upper limits of the fatigue damage which rainow analysis estimates at all the taps on a
side face of a square building. The third closed-form expression estimates larger fatigue
damage than the upper limits of the fatigue damage which rainow analysis estimates at all
the taps except S2. In S2, the difference between the upper limits of the fatigue damage
which rainow analysis estimates and the fatigue damage which the third closed-form
expression estimates is very small. The fourth closed-form expression estimates larger
fatigue damage than the upper limits of the fatigue damage which rainow analysis
estimates at all the taps except S19, S21, S22, and S23. In S19, S21, S22, and S23, the
difference between the upper limits of the fatigue damage which rainow analysis estimates
and the fatigue damage which the fourth closed-form expression estimates is very small.
ARTICLE IN PRESS
Table 3
Cases that D
cfe
is smaller than upper limits of D
rf
Tap no. D
cfe
( 10
4
) The upper limits of D
rf
( 10
4
)
S2 D
3rd, cfe
0.2056 0.2088
S19 D
4th, cfe
0.3307 0.3384
S21 D
4th, cfe
0.1781 0.1820
S22 D
4th, cfe
0.2142 0.2178
S23 D
4th, cfe
0.0164 0.0176
Note: D
cfe
is the fatigue damage estimated by the closedform expression, D
3rd, cfe
the fatigue damage estimated
by the third closed-form expression, D
4th, cfe
the fatigue damage estimated by the fourth closed-form expression
and D
rf
the fatigue damage estimated by rainow analysis.
N.-H. Ko, Y.-M. Kim / J. Wind Eng. Ind. Aerodyn. 95 (2007) 119 17
Therefore, the closed-form solutions that avoid the extensive numerical calculations are
applicable to estimate the wind-induced high-cycle fatigue damage in fasteners of curtain
wall on a side face of a square building for initial design calculations.
Acknowledgements
This work was supported by NRL (National Research Laboratory) Program of the
Ministry of Science and Technology, Republic of Korea.
References
[1] N.E. Dowling, Mechanical Behavior of Materials, Prentice-Hall, Englewood Cliffs NJ, 1999.
[2] J.W. Miles, On structural fatigue under random loading, J. Aeronaut. Sci. 21 (1954) 753762.
[3] S. Sarkani, D.P. Kihl, J.E. Beach, Fatigue of welded cruciforms subjected to narrow-band loadings, J. Eng.
Mech. 118 (2) (1992) 296311.
[4] A.G. Davenport, The estimation of load repetitions on structures with application to wind induced fatigue
and overload, RILEM International Symposium on the Effects of Repeated Loading of Materials and
Structures, Mexico City, Mexico, 1966.
[5] K. Patel, P. Freathy, A simplied method for assessing wind-induced fatigue damage, Eng. Struct. 6 (1984)
268273.
[6] T.A. Wyatt, An assessment of the sensitivity of lattice towers to fatigue induced by wind gusts, Eng. Struct. 6
(1984) 262267.
[7] B.A. Lynn, T. Stathopoulos, Wind-induced fatigue on low metal buildings, J. Struct. Eng. ASCE 111 (4)
(1985) 826839.
[8] S.R. Winterstein, Non-normal responses and fatigue damage, J. Struct. Eng. 111 (10) (1985) 12911295.
[9] S.R. Winterstein, Nonlinear vibration models for extremes and fatigue, J. Eng. Mech. ASCE 114 (10) (1988)
17721790.
[10] H.J. Gerhardt, C. Kramer, Wind induced loading cycle and fatigue testing of light weight roong xations,
J. Wind Eng. Ind. Aerodyn. 23 (1986) 237247.
[11] Y.L. Xu, Wind-induced fatigue loading on roof cladding of low-rise buildings, Technical Report No. 41,
Cyclone Testing Station, Department of Civil and Systems Engineering, James Cook University, Townsville,
Australia, 1993.
[12] Y.L. Xu, Fatigue performance of screw-fastened light-gauge-steel roong sheets, J. Struct. Eng. ASCE 121
(3) (1995) 389398.
[13] Y.L. Xu, Fatigue damage estimation of metal roof cladding subject to wind loading, J. Wind Eng. Ind.
Aerodyn. 72 (1997) 379388.
[14] C.W. Letchford, H.S. Norville, Wind pressure loading cycles for wall cladding during hurricanes, J. Wind
Eng. Ind. Aerodyn. 53 (1994) 189206.
[15] E.D. Jancauskas, M. Mahendran, G.R. Walker, Computer simulation of the fatigue behaviour of roof
cladding during the passage of a tropical cyclone, J. Wind Eng. Ind. Aerodyn. 51 (1994) 215227.
[16] M. Mahendran, Steel roof claddings under simulated cyclonic wind forces, Civil Eng. Trans. IEA, CE 36 (1)
(1994) 110.
[17] M. Mahendran, Wind-resistant low-rise buildings in the tropics, J. Perform. Constructed Facilities, ASCE, 9
(4) (1995) 330346.
[18] M.A. Mikitarenko, A.V. Perelmuter, Safe fatigue life of steel towers under the action of wind vibrations,
J. Wind Eng. Ind. Aerodyn. 74-76 (1998) 10911100.
[19] J.D. Holmes, Fatigue life under along-wind loading closed-form solutions, Eng. Struct. 24 (2002) 109114.
[20] M.P. Repetto, G. Solari, Dynamic alongwind fatigue of slender structures, Eng. Struct. 23 (2001) 16221633.
[21] M.P. Repetto, G. Solari, Dynamic crosswind fatigue of slender vertical structures, Wind Struct 5 (2002)
527542.
[22] M.P. Repetto, G. Solari, Directional wind-induced fatigue of slender vertical structures, J. Struct. Eng. 130
(7) (2004) 10321040.
[23] T.A. Wyatt, Determination of gust action stress cycle counts for fatigue checking of line-like steel structures,
J. Wind Eng. Ind. Aerodyn. 92 (2004) 359374.
ARTICLE IN PRESS
N.-H. Ko, Y.-M. Kim / J. Wind Eng. Ind. Aerodyn. 95 (2007) 119 18
[24] M.A. Miner, Cumulative damage in fatigue, Trans. ASME, 67 (1945) A159A164.
[25] The committee on Fatigue and Fracture Reliability of the Committee on Structural Safety and Reliability of
the Structural Division, Fatigue reliability: Variable amplitude loading, J. Struct. Div. ASCE, 108 (ST1)
(1982) 4769.
[26] L.D. Lutes, S. Sarkani, Stochastic analysis of Structural and Mechanical Vibrations, Prentice-Hall,
Englewood Cliffs, NJ, 1997.
[27] N.E. Dowling, Fatigue failure predictions for complicated stress-strain histories, J. Mater. 7 (1) (1972) 7187.
[28] D.V. Nelson, Cumulative fatigue damage in metals, Ph.D Thesis, Stanford University at Stanford California,
1978.
[29] C. Amzallag, J.P. Gerey, J. Bahuaud, Standardization of the rainow counting method for fatigue analysis,
Fatigue 16 (1994) 287293.
[30] P.H. Wirsching, M.C. Light, Fatigue under wide band stresses, J. Struct. Divi. ASCE, 106 (ST7) (1980)
15931607.
[31] K. Ortiz, N.K. Chen, Fatigue damage prediction for stationary wide-band stresses, ICASP 5, Presented at the
Fifth International Conference on the Applications of Statistics and Probability in Civil Engineering,
Vancouver, Canada, 1987.
[32] H.L. Perng, Damage accumulation in random loads, Ph.D. Dissertation, University of Arizona, Tucson, AZ,
1989.
[33] L.D. Lutes, M. Corazao, S.J. Hua, J. Zimmerman, Stochastic fatigue damage accumulation, J. Struct. Eng.
110 (11) (1984) 25852601.
[34] Y.M. Kim, Pressure uctuations on a tall building with square cross-section, Proceedings of International
Symposium on Wind and Structures (Cheju, Korea, January, 2000), Techno-Press, 2000, 393-405.
[35] J.A. Peterka, J.E. Cermak, Wind pressures on buildings probability densities, J. Struct. Eng. ASCE, 101
(ST6) (1975) 12551267.
[36] K.R. Gurley, Modelling and simulation of non-Gaussian processes, Ph.D. Dissertation, University of Notre
Dame, 1997.
[37] K.R. Gurley, A. Kareem, Analysis interpretation modeling and simulation of unsteady wind and pressure
data, J. Wind Eng. Ind. Aerodyn. 6971 (1997) 657669.
[38] D.A. Reed, E. Simiu, Wind loading and strength of cladding glass, J. Struct. Eng. ASCE, 110 (4) (1984)
715729.
[39] ASCE Manual, Wind tunnel studies of buildings and structures, ASCE Manuals and Reports on Engineering
Practice No.67, 1999
[40] E. Ishizaki, Problems in designing window glass against wind pressure, Wind Effects on Structures in
Proceedings of the Second U.S.A.Japan Research Seminar on Wind Effects on Structures, University of
Tokyo Press, 1976, 219225.
[41] D.E. Allen, W.A. Dalgliesh, Dynamic wind loads and cladding design, Prelim. Public. Intl. Assoc. of Bridge
and Structural Eng. Symposium (Lisbon) (1973) 279285.
[42] P.H. Wirsching, A.M. Shehata, Fatigue under wide band random stresses using the rain-ow method, J. Eng.
Mater. Technol. ASME, 99 (3) (1977) 205211.
ARTICLE IN PRESS
N.-H. Ko, Y.-M. Kim / J. Wind Eng. Ind. Aerodyn. 95 (2007) 119 19

Вам также может понравиться