Vous êtes sur la page 1sur 12

Journal of Wind Engineering and Industrial Aerodynamics, 41-44 (1992) 229-240 229

Elsevier

The measured and predicted response of a 300 m concrete chimney

J.L.Waldeck

Engineering Structures Programme, Division of Building Technology, CSIR, P.O. Box


395, Pretoria 0001, South Africa

Abstract
Various comparisons between the observed across-wind response of a 300m concrete
chimney and the response predicted by the well known model developed by Vickery
and Ba~u [1983] are presented. It is shown that the model performs remarkably well
when the predicted response is based on the observed total damping, which suggests
that the aerodynamic damping is positive for the small amplitudes characteristic of
concrete chimneys.

1.Introduction

At least three models for predicting the across-wind response of chimneys have been
developed during the past decade [Viekery and Basu, 1983; ESDU, 1985; Ruscheweyh,
1988], which once again stressed the importance of full-scale validation. This paper
presents the results of a comparison between the observed response of a chimney and
the predicted response based on the simplified version of the model developed by
Vickery and Basu [1983]. This model was selected primarily on the grounds of its
widespread acceptance in codes of practice (CICIND, ISO, NBC).

2. Experimental procedure

The experiment involved the instrumentation of a 300 m concrete chimney at the


Duvha power station near Witbank in South Africa. Measurements of the approach
wind conditions, the aerodynamic loads on the wind shield as well as the response of
the chimney extended over a period of about four years. The aerodynamic loads were
derived from 48 pressure transducers installed at elevations of 207 m, 252 m and 293 m.
The free-stream dynamic pressure was obtained by means of an indirect metilod based
on the observed pressure distribution around the chimney [Waldeck, 1989]. Unlike
previous applicatioa~ of the method, no assumptions regarding the mean pressure

0167-6105D2/$05.00 © 1992 ElsevierSciencePublishers B.V. All rights reserved.


230

distribution were necessary because the pressure transducers were referenced to the
free-stream static pressure by means of a static pressure probe suspended from a cable
between two chimneys. Displacements were obtained by integrating orthogonal
accelerations measured at an elevation of 291 m, using a frequency-domain technique
which included a correction for the component of gravitational acceleration induced by
the tilting of the accelerometers. More detailed descriptions of the instrumentation
have been given previously [Waldeck, 1983; 1989].

3. Quantification of parameters
The model developed by Vickery and Basu [1983] represents the vortex-shedding
force as a narrow-band random process with a lift spectrum centered on the Strouhal
frequency and bandwidth dependent on the intensity of turbulence. The
motion-induced force is described by an amplitude dependent aerodynamic damping
term which is negative in the vicinity of the critical wind speed. The motion-induced
force is assumed to be uncorrelated with the vortex--shedding force. The detailed
version of the mode! [Basu and Vickery, 1983] also considers the effect of the lateral
component of turbulence.

3.1.RMS fir coefficient


The R M S liftcoefficientdue to vortex shedding alone, 0LV , was estimated by fitting
a modified Gaussian distribution [Vickery and Clark, 1972] to the normalized power
spectral density (PSD) of the observed lift coefficient,CL(t ). This coefficientwas
based on the liftforce obtained by integrating the instantaneous pressure distribution
around the chimney and a free--stream dynamic pressure obtained from the indirect
method. The modified Gaussian distributionis given by

nSL(n ) = kn
O~ f~Bns exp
I (I-n/ns)2]
- B~ (1)

with SL(n ) the PSD and 0 L the R M S of the observed liftcoefficient,ns the mean
vortex shedding frequency and B the bandwidth parameter. The constant k represents
the fractionof the overallvariance resultingfrom vortex shedding, that is :

CLv : (2)

Strictly speaking, this estimate of (~LValso conta;~ns a small contribution from the
lateral component of turbulence, which cannot be separated from the vortex shedding
contribution at wavelengths in the vicinity of D/S.
231

The unknowns k, ns and B were determined from a least squares fit of eqn.(1) to the
normalized PSD of SL(n) as illustrated in Figure 1. The estimates of k and (~LVin
Table 1 were obtained at 252 m but are assumed to be representative of the top third
of the chimney.

"c:)
.-.,:
GAUSSIAN DISTRIBUTION

~m
N,.I

C
ii
i.n
C

c::)_

DATASET : 4
z +: 252,/,
J
f
S : 0,22
B = 0.31
k : 0,63
I (l~l I I
io"
n IHz}

Figure 1 Example of observed lift coefficient approximated by Gaussian distribution

3.2.Strouhal number
Estimates of the Strouhal number, S = nsD/0, and the bandwidth parameter, B, at
252m are given in Table 1. The mean wind speed, 0, was derived from the
fundamental relationship between wind speed and dynamic pressure by neglecting the
small influence of free-stream turbulence [Waldeck, 1989].

3.3.Damping
The total damping was estimated from the response of the chimney to wind
excitation, using the autocorrelation method [Cherry and Brady, 1965; Jeary and
Winney, 1972] and the method of spectral moments [Vanmarcke, 1977]. The results of
the latter method are not presented here because the autocorrelation method produced
somewhat more consistent results. Well-behaved autocorrelations were obtained by
removing the low frequency components of the displacement using a high-pass filter
with a -3 dB frequency of 0,0? Hz.
232

Table 1 Experimental results

Data U' fi]O' k (~Lv 13 n,,), Ct, y Q- G,~ ~,, O/O,:,-it :/,.o
set (m/s) (Hz) (/~cri,) (]~c,'it) (/~,:rit) (]~c,'it)

4 18,2 0,06 0,63 0,09 0,22 0,31 0,137 0,0082 0,0024 0,0010 0,0014 1,39 0,06
5 0;6 0,10 0,75 0,27 0,23 0,41 0,138 0,0050 0,0030 0,0005 0,0031 0,77 0
6 1~,1 0,04 0,73 0,08 0,17 0,56 0,139 0,0042 (0,00J6) 0,0009 (0,0007) 0,96 0,03
7 26,0 0,07 0,58 0,07 0,19 0,46 0,139 0,0054 0,0058 0,0015 0,0043 1,70 0
9 13,6 0,01 0,34 0,02 0,24 0,18 0,140 0,0040 0,0034 0,0008 0,0026 1,13 0,36
10 7,0 0,07 0,60 0,08 0,22 0,65 0,130 (0,0074)(0,0046) 0,0005 (0,0041) 0,62 0
13 22,1 0,05 0,65 0,07 0,23 0,32 0,137 0,0065 0,0062 0,0013 0,0040 1,78 0
14 17,7 0,03 0,53 0,06 0,17 0,50 0,137 (0,0062) 0,00~8 0,0010 (0,0028) 1,00 0,48
15 10,8 0,10 0,53 0,14 0,22 0,43 0,138 0,0046 0,0024 0,0006 0,0018 0,83 0
16 13,8 0,02 0,83 0,13 0,14 0,93 0,138 (0,0068) 0,0039 0,0008 0,0031 0,66 0
17 21,1 0,08 0,76 0,08 0,18 0,61 0,138 n/a nJa nJa nJa 1,33 0,06
18 20,7 0,05 0,55 0,06 0,22 0,39 0,138 0,0055 (0,0054) 0,0012 (0,0042) 1,64 0

Avcrage 0,138 0,0054 0,0041 0,0031


Coefficient of variation 0,009 0,25 0,34 0,37

® Total damping
The total along-wind damping, ~t, x, and across-wind damping, ~t, y, are presented
in Table 1 as a fraction of critical damping. Values for non-stationary data sets,
identified by means of the run test for stationarity [Bendat and Piersol, 1971], are
shown in brackets. These values were not used to compute the average and coefficient
of variation.

• Structural damping
The structural damping, (s, was determined by subtracting an estimate of the
aerodynamic damping from the total damping in the along-wind direction, (t, x. The
aerodynamic damping in the along-wind direction [Vickery and Kao, 1972] is given by

0 pDCD(z)0(z)¢l(zldz ] {(2~nl)' Mr} (3)

in which Ml is the mode-generalized mass and ~l the mode shape of the fundamental
mode. The calculated values of the aerodynamic damping and the corresponding
estimates of the structural damping, given by ~s = ~t, x-~a,x, are given in Table ).
These are based on Ml = 7 823 460 kg, nt = 0,138 Hz, C D = 0,52 [Waldeck, 1990] and
the relatively unimportant assumption that U(z) can be represented by a power law
with an exponent of 0,2.
233

= Aerodynamicdamping
The total damping in the across-wind direction, ~t, y, is compared with the the total
damping in the along-wind direction, (a, x, in Figure 2. The curve markers for non-
-stationary data sets are enclosed in brackets.

,4,,i Iv)
k.n I=) / o D4.1
(=k =~ a DS.l
tn + O6.1
{ x x D7.1
o Dg.I
v D10.1
B DI3.1
B DI4.1
a D15.1
= DI6.1
• DI7.1
* DIS. l
¢;. !
I
0.000 0.005 0.010 0.015

Figure 2 Comparison of total damping in the along- and across-wind directions

Assuming for the moment that the structural damping is the same in both
directions, the higher total damping in the across-wind direction can only be explained
by a positive aerodynamic contribution in the across-wind direction. Based on the
observation that the highest RMS across-wind displacement (~/D ~ 0,002)
corresponded to the only data set (7) for which (t, y was less than (t, x, it is postulated
that the aerodynamic damping becomes negative only when a certain displacement is
exceeded. The same tendency of a higher total damping in the across-wind direction,
with the exception of the data set with the highest RMS displacement (~/D ~ 0,003),
has also been noted by Basu [1983] while analysing the full-scale measurements of
Christensen [1977], but this was attributed to higher RMS displacements in the
across-wind direction. The displacements for the Duvha exr.eriment were, however,
too small to support such a relationship between displacement and damping.

The notion of positive aerodynamic damping is in stark contrast with the model of
Basu and Vickery [1983] and will be reconsidered later.
234

4. Comparison of observed and predicted response

The simplified response prediction model [Vickery and Basu, 1983] is given by

(~LV PD3{ ~r~- Lc }1/2.


8r~S 2
H
me

¢2 (Z)dz
}1/2
2(H/D -t- 2)
~](s "}" Ca, y
~(B, O/Ocr )
(4)

with
(5 /0 ,.)3/2 (1- Ocr/O)2
¢(B, (l]Ucr) = ° }
B2

Ucr = nlD/S

Q,y = Ka(-pD~/me) (5)

The simplified model was developed from the detailed model [Basu and Vickery,
1983] by introducing the following assumptions •

e The wind speed is constant and equal to the average over the top third of the
chimney
• The contribution to the lift force by lateral turbulence is negligible compared to the
vortex-shedding contribution
® The response remains within the small amplitude region where the total damping is
linear and positive, at a value below that provided by the structure.

The model was validated by comparing the observed RMS displacement of the
chimney with the predicted displacement based on the observed values of (~LV, S, B, nl
and Ct, y = ¢s + ¢a,y given in Table 1 and a correlation length (Lc ~ 1,5) computed
from the observed narrow-band cross-correlation of the lift force. The computed
fundamental mode shape, ~1, could not be verified experimentally but owing to the near
uniform cross section of the chimney, this had a minor influence on the equivalent
mass, me ( ~ 104500 kg]m).

The results of the compar!mn are presented in Figure 3. The average ratio between
the observed and predicted response is 1,02 and the coefficient of variation 0,25. Note
that the lowest predictions, for data sets 4, 13 and 18, are associated with mean wind
speeds well beyond the critical wind speed, which would not be considered in a typical
235

i=

"0
II
& •
REC
"O o 0~.1
+ A 05.1
U
It.
Q. + 06.1
0 B x 07.1
I>, la I
o D9. I
v D10.1
B B D13. I
m Of 4.1
D 015. ]
x D16. !
• 017. X
• BiB. I
! I I
0.00 0.02 0.0t O. 06 O. OB

Y"(observed) (m]

Figure 3 Comparison of observed and predicted RMS displacement (observed total


damping)

design situation. The coefficient of variation is not unreasonable in view of the


inherent variability in estimates of the total damping and the effect of a finite record
length on estimates of the RMS response. Vickery and Basu [1984] estimated the
coefficient of variation of estimates of the RMS displacement of a lightly d~tmged
system to random excitation as follows •

C of v --- 1/J 8~tnzT (6)

where T is the duration of the recording in seconds. Substituting (t ~ 0,005, nl - 0,138


Hz and T = 2130 s this equation predicts a coefficient of variation due to the finite
record length alone of 0,29.

The influence of aerodynamic damping is illustratedby repeating the comparison


between the observed and predicted response, but substituting the sum of the observed
structural damping (Table 1) and the aerodynamic damping suggested by Basu and
Vickery [1983] for the observed total damping. The value of the aerodynamic damping
is obtained from Figure 4, using Kao, max "- 0,9 [Basu and Vickery, 1983], where Kao is
236

the aerodynamic damping parameter for small amplitudes. The relationship between
Ka ( ~ Kao due to small displacements ) and the aerodynamic damping as a fraction of
critical damping is given in eqn.(5). The wind speed ratio, V]Ucrit, and the intensity
of turbulence for each data set, Iu, derived from the fluctuating pressure near the
stagnation point [Waldeck, 1990] are given in Table 1.

1,0 0
.... 0,1
..... 0,2
..... 0,3
0,8

z. 0,6
o
11
v

v g 0,4

. . . .

0,2
,/." .I
fJ ~

0,8 1 1,2 1,4 1,6


m

U/Ucrit

Figure 4 Influence of large-scale turbulence on Kao/Kao, max [Redrawn from Basu and
Vickery, 1883]

The results of the comparison are shown in Figure 5. Note that the most obvious
difference between the comparisons in Figures 3 and 5 correspond to data sets (4, 9, 14
and 17) with negative aerodynamic damping (Kao > 0) in Table 1. The fact that the
predicted response for these data sets is invariably higher than the predicted response
based on the observed total damping once again suggests that the aerodynamic
damping is positive for small amplitudes.

5.Practical application of the simplified model

The previous section dealt with the validation of the model by comparing the
observed response with the predicted response based on observed aerodynamic
parameters. To investigate the value of the model in a practical situation where these
237

mm X

A 0
=o
QI
W
"o
Qt
¢.

o.oo
I
0.02 0:04 o o6
i o.oa

(observed) (m)

Figure 5 Comparison of observed and predicted RMS displacement (suggested


aerodynamic damping)

parameters have to assessed from the literature, the following comparison is based on
typical input parameters which might have been used for the design of the Duvha
chimney. The mean wind speed is based on the measuremerts at the 252 m level and
the other parameters are assumed to have the following values :

(~LV= 0,18 Vickery and Basu [1984]


Kao = 0,47 Vickery and Basu [1984]
S - 0,19 Vickery and Basu [1984]
B = 1 Vickery and Basu [1984]
Lc - 1 Vickery and Basu [1984]
(s = 0,006 Average of full-scale measurements, Waldeck [1990]
nl - 0,15 Hz Computed, Waldeck [1990]

The results of the comparison are shown in Figure 6. The low predicted
displacements for data sets 5, 10 and 16 can be explained by very low values of ~(B.
0 / 0 c r i 0 obtained for 0 < < 0crit. These are not significant because the designer is
only concerned with the maximum response which occurs at 0 = 1,10,:tit. What is of
238

concern is the degree of conservatism in the predicted response of those data sets (6, 9
and 14) for which 0 = Ocrit. The two most important reasons for this observation are
the low overall damping resulting from the assumed negative aerodynamic contribution
and the much higher value of CLV ( = 0,18) compared to the average observed value of
only C L v - 0,1. This is demonstrated in Figure 7 by another comparison using the
same parameters except (JLV= 0,1 [Waldeck, 1990] and Kao = 0. The results are
clearly much more satisfactory.

m 0

/ o o~.a I
,, / I
~. ~l o / + o6.Z I
,~" / x 07.! !.

i O, O0 O,02 O,O~ O.06


!
O,08 O. I0

Vt0bserved) (m)

Figure 6 Comparison of observed and predicted RMS displacement based on typical


design parameters

6. Conclusion

The simplified model is validated by comparing the observed response of the


chimney with the predicted response based on observed aerodynamic and structural
parameters. The model is remarkably accurate considering that the comparisons
involve a large number of parameters with some uncertainty associated with each.

The predicted response based on recommended values of the RMS lift coefficient
and aerodynamic damping parameter tends to overestimate the response when the
computed total damping is less than the observed total damping, which can only be
explained by a positive aerodynamic damping term. The notion that the aerodynamic
239

damping is positive for small amplitudes is supported by the observation that the total
damping in the across-wind direction is higher than in the along-wind direction. This
represents a radical departure from the model developed by Vickery and Basu [1983]
but need not cause concern because such low amplitudes are not important from a
design point of view.

"0

.L_.
iJ /" o q4. !
/ A ~05.1
Q.
m ,o + D5 1

/ o o_9.1
/ ,.. v DlO. 1

zz m DI4.1
a DI5.
m DI6.1
• D17.1
* O18. I
D
A I I i
6.00 0.02 0.04 0.06 0.08 O. lO

"Y Iobserved} Ira)

Figure 7 Comparison of observed and predicted RMS displacement based on revised


design parameters

References
Basu, R.I. (1983) Across-wind response of slender structures of circular
cross-section to atmospheric turbulence. University of Western Ontario, London,
Ontario, Canada, BLWT-3-1983.
Basu, R.I. and Vickery, B.J. (1983) Across wind vibrations of structures of
circular cross-section. Part II. Development of a mathematical model for
full-scale applications. Journal of Wind Engineerin9 and Industrial Aerodynamics,
Vol. 12, pp. 75-97.
Bendat, J.S. and Piersol, A.G. ( 1 9 7 1 ) Random data : analysis and
measurement procedures. Wiley-Interscience.
Cherry, S. and Brady, A.G. (1965) Determination of structural damping
properties by statistical analysis of random vibrations. Proceedings of the third
world conference on earthquake engineering, New Zealand.
Christensen, O. (1977) Wind forces on and excitation of a 130m concrete
chimney. Series Paper No.15, Institute of Hydrodynamics and Hydraulic
Engineering, Technical University of Denmark.
240

Engineering Sciences Data Unit. ( 1 9 8 0 ) Mean forces, pressures and flow


field velocities for circular cylindrical structures: Single cylinder with
two--dimensional flow. Item No. 80025, ESDU International, London.
Engineering Sciences Data Unit. (1981) Mean forces, pressures and
moments for circular cylindrical structures: Finite length cylinders in uniform and
shear flow. Item No. 81017, ESDU International, London.
Engineering Sciences Data Unit. ( 1 9 8 5 ) Circular--cylindrical structures:
dynamic response to vortex shedding. Part I: Calculation procedures and
derivation. Item No. 85038, ESDU International, London.
Jeary, A.P and Winney, P.E. (1972) Determination of structural damping
of a large multi-flue chimney from the response to wind excitation. Proceedings
of the Institution of Civil Engineers, Vol. 53, London, pp. 569-577
Ruscheweyh, H. (1988) Cross wind vibration of steel stael~s. A critical
comparison of various code provisions. Sixth International Chimney Conference,
Brighton, England.
Vanmarcke, E.H. (1977) Method of spectral moments to estimate
structural damping. Stochastic problems in dynamics, edited by Clarkson, B.I.,
Pitman, pp. 515-524.
Vickery, B.J. and Basu R.I. (1983) Simplified appro.~ches to the evaluation
of the across-wind response of chimneys. Journal of Wind Engineering and
Industrial Aerodynamics, Vol. 14, pp. 153-166.
Vickery, B.3. and Basu, R.I. (1.q84) The response of reinforced concrete
chimneys to vortex shedding. EngineeringStructures, Vol. 6, pp. 324-333.
Vickery, B.J. and Clark, A.W, (1972) Lift or across-wind response of
tapered stacks. Journal of the Structural Division, Proceedings of the ASCE, No.
ST1.
Vickery, B.J. and Kao, K.tt. (1972) Drag or along-wind response of slender
structures. Journal of the Structural Division, Proceedings of the ASCE, Vol. 98,
No. ST1, pp. 21-36.
Waldeck, J.L. (1983) h digital system for the measurement of wind effects
on large str~=ctures. Journal of Wind Engineerintt and Industrial Aerodynamics,
Vol. 13, pp. 453--464.
Waldeck, J.L. ( 1 9 8 9 ) Measurement of wind effects on a 300 m concrete
chimney - remrence parameters. Journal of Wind Engineering and Industrial
Aerodynamics, Vol. 32, pp. 109-210.
Waldeck, J.L (1990) Measurement of wind effects on a 300 m concrete chimney:
Final report. Internal report 90]02, Engineering Structures Programme, Division
of Building Technology, CSIR.

Vous aimerez peut-être aussi