Ashcroft 1994

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

dOt,~NAI.

OF

II Journal of Wind Engineering


ELSEVIER and Industrial Aerodynamics 53 (1994) 331-355

The relationship between the gust ratio,


terrain roughness, gust duration
and the hourly mean wind speed
J. A s h c r o f t
Scientific Support Group, The Meteorological O~ce, Johnson House,
London Road, Bracknell, Berkshire RGI2 2SY, UK

Received 14 May 1993; accepted 15 March 1994

Abstract

A large volume of 1 min mean wind data is available at the United Kingdom Meteorological
Office. Until recently, very little scientific use has been made of the data. This study provides
up-to-date information, using this data, upon the relationships between the gust ratio, the
terrain roughness and the hourly mean wind speed by gust duration. The ratio of the maximum
wind averaged over a period of a few seconds, 1 and 10 min and the hourly mean is obtained
after careful quality control of many hours of data for each of eight wind direction sectors at 14
Meteorological Office anemograph stations. The ratio of the peak recorded gust to the 10 rain
mean wind is also derived using two alternative derivations of the 10 min mean wind. The
sector median gust ratio is successfully correlated with the estimate of terrain roughness derived
from the best estimate of the sector aerodynamic roughness length. The form of the gust
ratio-terrain roughness dependence is very similar to previous relationships for gust averaging
periods ranging from a few seconds to 10 min but the median gust ratios obtained for averaging
periods of 1 and 10 min are somewhat lower than those given in previous guidance documents.
The ratio of the 3 s gust and the 10 min mean wind is obtained. A comparison is made between
the observed median values of the above gust ratio and various predictive relationships in the
literature. The dependence of the median gust ratio upon hourly mean wind speed is obtained.
The median gust ratios for gust averaging periods of 1 and 10 min averaged over the set of
stations do show a statistically significant decline with increasing wind speed. The 3 s gust ratio
shows a much wider range of wind speed dependence from station to station and it is not
possible to confidently define a general pattern of change with wind speed.

Nomenclature

d zero p l a n e d i s p l a c e m e n t
dir6o 1 min w i n d direction

Elsevier Science B.V.


SSD! 0 1 6 7 - 6 1 0 5 ( 9 4 ) 0 0 0 0 9 - 3
332 J. Ashcroft/J. Wind Eng. Ind. Aerodyn. 53 (1994) 331-355

dirm mode of 1 min wind directions in the hour


G6oo ratio of maximum of 10 min mean winds (non-overlapping average)
to hourly mean
G6o0c ratio of maximum of 10 min mean winds (moving average) to hourly
mean
660 ratio of maximum of 1 min mean winds to hourly mean
Gp ratio of peak recorded gust to hourly mean
6~ ratio of peak recorded gust to 10 min mean
(non-centred mean)
6~o ratio of peak recorded gust to 10 min mean (centred mean)
Gpo ratio of peak recorded gust to 10 min mean estimated from Gp and
6600
lu turbulent intensity - the rms of wind speed fluctuations as a propor-
tion of the hourly mean wind
Kr non-dimensional terrain roughness parameter
l gust wavelength
L maximum likely gust wavelength
n6o number of 1 min winds (non-missing) in the hour
RMSD root mean square of 1 min wind directions about the modal wind
direction
temporal change in 1 min winds over the hour relative to the hourly
mean wind
t averaging duration of the gust
T duration over which the wind is sampled
U600 10 min mean wind (non-overlapping average)
U600c 10 min mean wind (moving average)
U60 1 min mean wind
Up peak recorded gust speed
hourly mean wind speed
U' the ratio In [(z - d ) / Z o ]
Z height above ground level
Z0 aerodynamic roughness length
Zg gradient wind height
~600,3600 coefficient of variation of the 10 min mean, i.e. the ratio of the
standard deviation of the 10 min wind (non-overlapping average)
about the hourly mean as a proportion of the hourly mean wind

1. Introduction

The Digital Anemograph Logging Equipment (DALE) is a data logging system


which works in conjunction with the Meteorological Office Mark 4 or Mark 5 anemo-
meter system [1]. The DALE has been set up to average the wind speed and direction
over sequential 1 min periods and record this information on magnetic tape. At the
end of every clock hour the maximum recorded gust, speed, direction and time are
also placed onto the magnetic tape [1]. The I min winds and 3 s gust data (also called
J. Ashcroft/J. Wind Eng. Ind. Aerodyn. 53 (1994) 331-355 333

here the peak recorded gust data) have been recorded at a number of anemo-
graph stations (mostly at civil or RAF airfields) for a number of years and this data
has been processed at the Meteorological Office, Bracknell to provide the hourly
mean wind speed and modal wind direction and gust information for the Me-
teorological Office's climatological anemograph data sets. The 1 min wind data has
been archived on magnetic tape since the autumn of 1983 but very little use has been
made of this data for answering enquiries or for the purpose of scientific research.
Durst [2] indirectly derived relationships between the likely maximum expected
gust (averaging periods of 0.5 to 600 s) and the hourly mean wind using wind
observations at Cardington. These relationships were incorporated with results from
Sale, Australia by Deacon [3] and have been summarized, along with the data of
Shellard in Ref. [4]. There is little other summarized data available for engineers on
the dependence of the gust ratio upon surface roughness. The work of Weiringa [5]
contains a set of gust ratio data from various sources along with a set of data for a site
in the Netherlands. The British Standards document BS8100 [6] usefully summarizes
the available information on the relationship between the gust ratio and the terrain
type for gust periods ranging from 1 to 600 s in both tabular and graphical form.
Ref. [7] contains much valuable information on the expected dependence of short
period gust ratios (1 to 12 s averages) upon terrain type and height above ground.
A number of other studies, see, e.g., Refs. [8,9] have derived relationships for the
probable maximum short period gust within the hour by integrating the power
spectrum of the horizontal wind velocity.
The large amount of DALE wind data available can, however, provide the neces-
sary information to define empirical relationships between the maximum of a set of
short period mean winds and the hourly mean wind and the dependence of the gust
ratios upon the local site roughness can be studied in a consistent way. It is the aim of
this work to provide information on these relationships using the DALE 1 min winds
and the peak recorded gusts. Unfortunately, no recorded wind data from the DALE
system is available over an averaging period between the few seconds assigned to the
recorded peak gust and the 1 min mean wind. A 15 s averaging period for the DALE
equipment would have been ideal, as gust ratio relationships for this averaging period
would have been most useful for engineers designing larger structures, and mean
winds over longer averaging periods could have been derived from these averages. In
the absence of such data, we can only consider the gust ratios derived from the peak
recorded gust data and winds averaged over periods of 1 rain or more.

2. Method of analysis of DALE wind data

2.1. Selection and quality control o f D A L E 1 min mean winds

The analysis was limited to the years 1988 to 1990 because of the very large volumes
of 1 min wind data available and restrictions of resources. The calendar year 1990 was
chosen in particular because of the strong winds experienced in January and February
of that year in southern England. The location of all the currently operational DALE
stations and the chosen set is shown in Fig. 1.
334 J. Ashcroft/J. Wind Eng. Ind. Aerodyn. 53 (1994) 331-355

The choice of stations was made on the basis of site descriptions held at the
Meteorological Office and a study of the land use around the anemograph site using
1:50 000 scale Ordnance Survey maps. The intention was to choose sites which were
surrounded by terrain of as wide a range of aerodynamic roughness as possible but,
unfortunately, at many DALE equipped sites the roughness is quite similar, since the
site consists of an airfield surrounded by flat or gently rolling country. Some DALE
stations were especially chosen because the site appeared to have an above average
roughness, for example, Peterhead and Gravesend anemograph stations are close to
urban and industrial areas and the Eskdalemuir site is heavily forested.

SUMBURGH

ESKDALEMUIR

,~nAUGHTON

.7 o j m/,._....~
/ " • • r'

f" LARKHILL •
l~i • GRAV~ND
NCEUX

Fig. 1. Location of currently operational DALE stations and stations used in the study.
J. Ashcroft/J. Wind Eng. Ind. Aerodyn. 53 (1994) 331-355 335

Other sites were chosen to provide information on gust ratios over smooth terrain.
The Shoeburyness site is very fiat and open in most directions looking away from the
anemometer mast, being mostly marsh fiats. The Chivenor site was chosen to provide
information upon the gust ratios for wind flow over estuarine mud and sand fiats. In
certain sectors the Sumburgh and Benbecula sites provide information on gust ratios
for onshore winds onto a low lying coast. The wind data for the DALE stations at
Herstmonceux, Cranwell and Coningsby was extracted to provide information on
gust factors over open countryside of varying roughness. The wind data from Heath-
row and Abbotsinch, it was hoped, would be useful for defining the gust ratios in
suburban areas.

2.2. Quality control of D A L E wind data

A number of routine checks were made upon the DALE data such as checks for
spurious wind speeds and directions - for more detailed information see Ref. [10].
Any hour with more than two 1 min mean wind speeds or directions queried by
quality control routines was rejected from the analysis. In addition, it was thought
wise to reject all hours where the hourly mean wind speed was less than 6 knots
because of the fact that a few knots of wind are required to start up the Mark 4 and
Mark 5 cup anemometers from rest [1].
Another important consideration was pointed out by Weiringa [5]. Unless careful
checks are made to ensure that the mean wind over the hour is not subject to
a significant temporal trend or to sharp changes in magnitude due to, for example
a frontal passage, then the gust ratio (peak short period mean relative to the sampling
duration mean) will tend to be overestimated. This is because the longer period
variations of the mean wind would tend to be included within the peak gust value.
Since the sequence of one hours' 1 rain DALE mean wind data describes the temporal
variation of the wind then some check can be made to ensure that the wind over the
hour is sufficiently steady before the gust ratios are derived.
The procedure adopted was to regress, for each clock hour, the 1 min winds versus
time, i.e. the minute values within the hour. The slope of this linear regression over the
hour gives an estimate of the temporal change in the wind. This latter value, divided
by the hourly mean wind (calculated from the 1 min mean winds) must be small if the
mean wind is to be regarded as steady in time. The relative slope, S of the regression of
the 1 min winds to the hourly mean is defined as
S = slope of regression x 60/~. (1)
After studying a large number of hours of data at sites with varying roughness and at
varying wind speeds it was decided that the critical value for rejection of the hour's
data on temporal changes alone should be S > 0.25. This value would eliminate those
hours with the most serious trends in the wind but would not lead to an excessive
reduction in the total number of hours of data available for the analysis.
To eliminate hours with large wind direction variations within the set of 1 min
means a check was made on the rms of the 1 min wind directions about the modal
wind direction (taken from the Meteorological Office climatological anemograph
336 J. Ashcroft/J. Wind Eng. Ind. Aerodyn. 53 (1994) 331-355

data sets for the DALE station). The rms is given by

RMSD = x/Z(dir60 - dirm)2/n60 . (2)

The most extreme variations in inter-hourly wind direction were found to be elimi-
nated by setting RMSD > 20 ° as a rejection criterion.
Another problem, revealed by plotting a sample of time series of 1 min winds, is that
of large inter-hourly variation of wind over a period of tens of minutes without any
significant overall temporal variation from the start to the end of the clock hour. The
method developed to check this was to calculate the standard deviation of the 10 min
winds about the hourly mean wind. The 10 min mean winds were obtained by
averaging the DALE 1 min winds in six successive and independent 10 min intervals
from the start to the end of the clock hour. The standard deviation of the six
independently derived 10 min winds about the hourly mean as a proportion of the
hourly mean value was calculated,

N/~"~(U600 -- fi)2/(n6o 0 -- 1)
?]600,3600 = (3)

It is expected that this coefficient of variation would depend upon the local terrain
roughness and also upon boundary layer stability, so it was not possible to pre-
determine a critical value for all sites and atmospheric conditions which, when
exceeded, would confidently indicate unsteadiness of the mean wind. As part of the
study of the variability of the 10 min mean winds at various sites [10] a number of
tests were undertaken on a critical value of the coefficient of variation in order to
determine acceptable values for this parameter for use in a check on the stationarity of
the wind. The final value chosen was 0.15 and was a compromise, reflecting the need
to check the stationarity of the wind and the need to avoid an excessive loss of data,
bearing in mind the short period of data available for the study.

2.3. Calculation of gust ratios

Once the stationarity check on the hourly mean wind was performed then the
calculation of the gust ratios could take place. The large volume of available DALE
data gives us the opportunity to directly calculate a large number of gust ratios and
directly derive representative estimates from their distributions. The practice in much
previous work has been to follow the procedure of Durst [2] and indirectly derive the
relationship between the peak short period average wind and the 10 min or hourly
mean wind by using an estimate of the standard deviation of the short period average
about the longer sampling period mean and a multiplier of this which would be
sufficient to define an extreme short period gust. This procedure was also followed by
Weiringa [5]. In addition, in Ref. [ 10] a number of estimates of the gust ratio (peak 10
min to the hourly mean) are also made in the manner of Durst [2] using values of
?]600,3600 derived from DALE data and the values obtained were found to be in fair
agreement with those obtained by Durst.
J. Ashcroft/J. Wind Eng. Ind. Aerodyn. 53 (1994) 331-355 337

The directly derived gust ratios are defined as follows: peak recorded gust (aver-
aging period of a few seconds) and the hourly mean wind over the clock hour define
the peak gust ratio

Gp = Up/~. (4)

The maximum 1 min mean wind in the given clock hour is easily found and with the
hourly mean wind for the same hour is used in the calculation of the 1 min gust ratio,

G6o ~- U6omax/U. (5)

The 10 min mean winds were previously calculated for use in the stationarity check
and the maximum of these six consecutive values within the hour to the hourly mean
wind defines the ratio

G6o 0 = U6o 0 max/U. (6)

It is possible to average the 1 min winds to form the 10 min mean in a number of ways.
The above process of calculation of the 10 min mean is perfectly consistent with the
standard practice of averaging in non-overlapping and independent time intervals.
This averaging is usually performed over a much shorter time interval of a few seconds
or tens of seconds to calculate short period mean winds within usually, a total time
period of 10 min at most - see the work of Durst [2] and Deacon [3] for example.
An alternative estimate of the 10 min mean wind may be obtained by a 10 min
moving average running through the sequence of 1 min mean winds. Because of
operational problems at the DALE stations and removal of data by quality control
checks the DALE data used in the study is not a sequential and chronological set of
hourly 1 min wind observations and it is not generally possible to form a moving
average for a given hour which involves the use of the subsequent and preceding
hour's 1 min wind data in the average. In order to use only the current hour's 1 min
winds and still define a 10 min moving average for each minute through the hour, the
set of 1 min values was artificially extended by 5 min before and after the hour using
the mean wind of the first 5 or last 5 min of true data so that the 10 minute mean wind
could be defined for each 1 min value (5 min before the given minute and 4 min after
define the 10 min period). The maximum of this sequence of 10 min means relative to
the hourly mean wind defines the alternative gust ratio G600c as

G6ooc : U6oo . . . . /1~. (7)

In addition, the ratio of the peak recorded gust to the 10 min mean wind was
calculated. The correct 10 min mean was found by referring to the recorded time of the
peak gust. Both non-centred and centred 10 min means were used,
t
Gp ~- Up/U6oo, (8)
!
Gpc : Ul~U6ooc. (9)
338 J. Ashcroft/J. Wind Eng. Ind. Aerodyn. 53 (1994) 331-355

All the above gust ratios were calculated for 8 direction sectors at each of the
14 stations. A finer division of the compass was not thought feasible due to the need to
have a sufficient sample size for every sector and wind speed class. The hourly modal
direction, obtainable from the Meteorological Office climatological anemograph data
sets, was used to define the correct sector for those gust ratios based upon the hourly
mean wind. The direction of the 10 min mean wind used in the calculation of G~ and
G~,c was taken as the mode of the directions of the 1 min mean winds in the specified
10 min period. The gust ratios were further classified by the hourly mean wind speed
class, namely: 6-10, 11-15, 16-20, 21-25, 26-30 and > 30 knot.
The gust ratios were calculated and placed in frequency distributions by ratio class,
direction class and wind speed class. F r o m these distributions the mean, mode,
median and other quartiles and the standard deviation were calculated. The indi-
vidual probability distributions were summed by selected wind speed classes to
provide a directional gust ratio probability distribution or for all or selected wind
direction sectors for selected wind speed classes to provide a distribution by wind
speed alone.

3. The dependence of gust ratios upon averaging time and terrain roughness

3.1. The probability distribution of the gust ratio

The frequency and probability distribution of the gust ratios G6oo, G6o and Gp,
derived from the D A L E data at the Shoeburyness and Eskdalemuir sites, was studied
as a number of stationarity checks were made on the hourly mean wind. There were
four different stationarity checks made.
Reject hour's data as non-steady if:
1. S > 0.25 or R M S D > 20 o or (S < 0.25 and ?~600,3600 > 0.15);
2. S > 0.25 or R M S D > 20 ° or (S < 0.25 and ~/600.3600 > 0.10);
3. S > 0.25 or R M S D > 20 °;
4. S > 0.10 or R M S D > 20°.
The purpose of undertaking a number of alternative checks was to ascertain the
sensitivity of the distributions of the gust ratios and their median values to any check
made on the stationarity of the wind and the likely loss of observations when more
stringent checks were made on the hourly change in mean wind, S and the limiting
value of ?]600, 3600"
Some examples of the probability distributions for the gust ratios G60o, G6o and
Gp obtained at the Eskdalemuir and Shoeburyness sites (all sector summation) over
a wide range of hourly mean speeds are shown in Figs. 2-4. Note that the gust ratio
distributions obtained for all occasions when the hourly mean speed was > 6 knot
will include cases when the boundary layer stability will be non-neutral. The effect of
an introduction of a stationarity check when occasions of low wind speeds are
included is to clearly reduce the proportion of values in the upper tail of the
distribution and to sharpen the peak of the distribution, particularly for 1 and 10 min
averages. There appears to be little difference between the distributions of G600 and
J. Ashcroft/J. Wind Eng. lnd. Aerodyn. 53 (1994) 331-355 339

0.6 ProporBon Key:


jJt type 0 check
#t
e t ------ typelcheck
0.5 t
, t ..... type 2 check
e t
I t ...... type 3 check
k ........... type 4 check
0.4. I t
, ,~ ...... .~,..
: r,.-...~.i,..~
e
0.3.

0.2.

0.1 4/
i ! ) i | ) ) ) ) ,
1.01-1.05 1.06-1.10 1.11-1.15 1.16-1.20 1.21-1.25 1.26-1.30 1.31-1.35 1.36-1.40 1,41-1.45 146-1.50

G600

Fig. 2. Probability distribution of the gust ratio G6oo at the Eskdalemuir site as a function of the
stationarity check.

0.30 Proportion Key:


type 0 check

E
~t ------ typeI check
) t ..... type 2 check
0.25 ' ,=.-. t
r-;~,~, ...... type 3 check
ii - ~ ........... type 4 check

020 "Ir--l
0.15

0.10

0.05

1.06 1.11 1.16 1.21 1.26 1.31 1.36 1.41 1.46 1.51 1.56 1.61 1.66 1.71 1.76 1.81 1.86
1.10 1.15 1.20 1.25 1.30 1.35 1.40 1.45 1.50 1.55 1.60 1.65 1.70 1.75 1.80 1.85 1.90

G60

Fig. 3. Probability distribution of the gust ratio G6o at the Shoeburyness site as a function of the
340 J. Ashcroft/J. Wind Eng. Ind. Aerodyn. 53 (1994) 331-355

0.15] Proportion Key:

aI - - type 0 check
0.14 t a a'~" type 1 check
0.13 ~ ~l
..... type 2 check
o.1;Zl me
" /
2,,.
• %
...... type 3 check
........... type 4 check
°.1' 1 .. ,,.';............
010 -e , /.:" .i"...\
o.o9-1
/ : I.,/ / ~
,'"~:" "~d-. ~.k
",..~:~'~-.
o.oeI ,~: / ~ ~,
o.o7-t ,'//: / ',~,
I ,~;y! I ,~. x
o.o64 :/~, I
e3~l
'.~.~.
#, I ',,",:,,,\
oq
001
0.021 ,"'t / ~'. % ".\
I * "~-"¢.~1 "" ° "". "~ ,.....~.~--'----~..

1.51 1.56 1.61 1.66 1. 2.71


1.55 1.60 1.65 1.70 1.75 1.84) 1.85 1.90 1.95 2.00 2.05 2.10 2.15 2.20 2.25 2.30 2.35 2.40 2.45 2.50 2.00 2.70 2.80

Gp

Fig. 4. Probability distribution of the gust ratio Gpat the Eskdalemuir site as a function of the stationarity
check.

G6o for checks 1 and 3 though the distribution of G6oois directly affected over rough
terrain, with a notable loss of the upper tail of the distribution of G6oo at the
Eskdalemuir site and a very severe loss of the original distribution when check 2 is
applied. This loss of data was considered unacceptable as it is obviously impinging on
the usual distribution of %00, 3600 appropriate to this terrain type. Check 4 gives
a very similar overall probability distribution of peak gusts to that of checks 1 and
3 though it was not adopted for subsequent general use as the loss of observations was
considered unacceptable. Thus checks 1 and 3 were considered suitable for use over
a wide range of wind speeds.
Check 1 was adopted for use in the generation of the gust ratio distributions at all
sites as it incorporates some check on the temporal trend in the wind over the hour
and any extreme inter-hourly wind speed variations due to longer period (and
probably) larger scale lulls and gusts. It is worthwhile noting that most of the loss of
data between checks 0 (all data) and check 1 is in fact achieved by the limit on the
slope, S and not by the limit on the value of %oo, 3600. This can be seen by comparing
the distributions for checks 1 and 3. Therefore the difficulty in deciding upon the
limiting value of r/is not in practice of much significance at the vast majority of sites.
The median value of the gust ratio which can be derived from the gust ratio
distribution is only affected by a seemingly small amount by the introduction of
a stationarity check on the wind speed. Using all occasions when the hourly mean
wind speed was 11 knot or more (approximately neutral conditions) a roughly 3%
reduction in the median peak gust was obtained when the type 3 stationarity check
J. Ashcroft/J. Wind Eng. Ind. Aerodyn. 53 (1994) 331-355 341

was performed upon the the DALE 1 min winds, compared to those obtained when no
stationarity checks were performed. However, the variation in the gust ratio Gp, for
example, over the majority of terrain types over land (according to Ref. [6]) is only
1.45-1.65 or 1.55 _+ 6.5% so the check is clearly of some importance in determining
the final gust ratio estimate from DALE data.
The distribution of gust ratios in occasions of high wind speeds (type 1 stationarity
check) for four stations with differing terrain roughness is shown in Figs. 5 and 6 for
G6o and Gp, respectively. The probability distribution of the gust ratio G6o shows
a modest dependence upon the differing site terrain roughness and has a very similar
form at all sites, the peak of the distribution being progressively shifted to the right as
the general terrain roughness increases. The distribution of the gust ratio Gp shows
a rather different shape, with a greater standard deviation and greater sensitivity to

0.4 Proportion Key:


- - , Eskdalemuir data

.t., -- -- -- Larkhill data

!\,' ..... Cranwell data


...... Shoeburyness data
r m~,l
I "!

0.3

I, I I,%

! i I 11/ 1 l

ill ~1, I
;I i ti: 1 I
;, i /i', ', l
0.2

"l!\i\
I,

g'l
I

I i.~ I l
i i, I '.',.; \
0.1
:, ~ I '.. ", ',
!l ~ I '._ ', ', %
/i: I ',',3, \
]:I II
;/~
"'k \
~. 1, .\
;, # I ~. ~',. "x
0.0 li''l /il~l ~I" I I I I I i ~ I I i----| I _ I._ I_ I
1.01 1.06 1.11 1.16 1.2"I 1,26 1.31 1.36 1.41 1.46 1.51 1.5S 1.51 1.66 1,71 1.75
1.05 1.10 1.15 1.20 1.25 1.30 1.35 1.40 1.45 1.50 1.55 1.60 165 1.70 1.75 180

G60

Fig. 5. Probability distribution of the gust ratio G6o at high wind speeds.
342 J. Ashcroft/J. Wind Eng. Ind. Aerodyn. 53 (1994) 331 355

0.3 Prolx~,ion Key:


- - Eskdalemuir data
Larkhill data
,.', ..... Cranwell data
i I. ...... Shoeburyness data
i i.
i i.
i ~.
0.2
al 'l I
Ii • ~.
q ~t A
/\
0.1
t
/ /
'..',',,,,/ \
ff
il

i v"
/
/

i
x
;I x

I
I ""

[/

/- ,,
0.0 i I I i I i i i i i i i i i i i i i
1.01-1.10 1.11-1,20 1.21-1.30 1.31-1.40 1.41-1.50 1.51 1.56 1.61 1.66 1.71 1,76 1.81 1.86 1.81 1.96 2.01 2.062.11 2,16 2.21
1.55 1.60 1.65 1.70 1.75 1.80 1.85 1.90 1.95 2.00 2.05 2 . 1 0 2 . 1 5 2.20 2.25
Gp

Fig. 6. Probability distribution of the gust ratio Gp at high wind speeds.

local terrain roughness. The distributions are very similar in shape at Cranwell and
Shoeburyness (two sectors omitted where there were a significant number of build-
ings) but at Larkhill and Eskdalemuir the distributions are broader and are skewed
towards higher values of the gust ratio. The much wider range of gust ratio at the
Larkhill site than at the other sites reflects the considerable directional variation of
roughness at the station.
These distributions were derived for each wind speed class individually and these
results showed that if a sufficient sample was available at the station in question, at
hourly mean wind speeds above 20 knot, the introduction of stationarity checks 1-3
made very little impact on the distribution of the peak gust ratios and gust ratios
averaged over longer periods. This implies that gust ratios derived from occasions of
high mean wind speeds are effectively those derived in conditions of stationary winds.
Therefore, if the gust ratio is required from information which is restricted to the
hourly mean wind and the peak recorded gust (i.e. no stationarity check is possible)
the analysis should be confined to those occasions of strong winds, i.e. hourly mean
wind speed roughly 21 knot or greater (say 10 m s - i or greater) if the gust ratios are
not to be overestimates.
In this particular study it was not feasible to search through the whole period of
archived D A L E data at every station to find those hours when the wind speed met the
above suggested criterion. Because of the need to provide a sufficient body of
observations in each wind direction sector for correlation of the median gust ratio
with sector roughness it was found necessary to use the sector gust ratio distributions
when the hourly mean wind speed within that sector was > 11 knot. This lower limit
J. Ashcroft / J. Wind Eng. Ind. Aerodyn. 53 (1994) 331-355 343

to the hourly mean wind speed will unfortunately include some occasions when the
boundary layer may be classified as being slightly unstable (for example, in high
summer with clear skies). However, once a large body of thousands of hours of
observations is brought together using the three years worth of DALE data then these
cases can be expected to be only a small percentage of the total number and the
median value of the gust ratio will be scarcely affected by their inclusion. Weiringa
[11] suggested that gust ratios may be derived for near neutral conditions when the
hourly mean wind speed, fi is > 6 m s- 1. This lower limit on the mean hourly speed is
practically the same as that used here.

3.2. Dependence of the median gust ratio upon terrain roughness category

3.2.1. Estimation of aerodynamic roughness length


In order to derive relationships between the median peak gust ratio and station
sector roughness it is necessary to have some numerical estimate of roughness. In this
study the aerodynamic roughness length Zo, or parameters dependent upon it, are
correlated by sector with the median gust ratio derived from the probability distribu-
tion of the gust ratios. It is possible to derive an effective roughness length from the
median gust ratio Gp using the technique of Weiringa [-11]. However this estimate
would not be independent of the gust ratio itself, so this method is not used here. The
sector roughness length was estimated instead from an assessment of the land use
around the anemometer site using 1:50000 Ordnance Survey maps and the terrain
descriptions in Section C of Part 2 of Ref. [6]. These estimates were substantiated
using the roughness lengths around the DALE station calculated from digitised land
use data sets available in the Meteorological Office using the technique described in
Ref. [ 12]. These roughness lengths were calculated for 100 m by 100m pixels and averaged
out over radial distances of 300 m in each of the eight 45 ° sectors at the 14 stations.
A particular problem concerns the distance over which the roughness length should
be estimated or calculated. The roughness length can be quite variable with distance
within any sector and a different averaging distance will result in quite different
estimates of Zo. Weiringa [11] suggested that the gust ratios at a typical anemometer
height of 10 m will be influenced by terrain roughness out to a distance of roughly
3 km and will especially reflect the disturbance to the flow induced by the largest
roughness elements in the path of the wind. This guideline was followed here in
deriving the effective roughness length by sector from the Ordnance Survey maps. The
sector average roughness lengths derived from the digitised land use data were
subjectively weighted towards the higher of the values obtained for the individual
300 m sectors. The intention was to produce a value of z0 similar to that which might
be obtained by applying Weiringa's technique.
In the great majority of cases the subjectively assessed roughness agreed well with
that derived from an assessment of the numerically calculated sector roughness but in
one or two cases, for example at the Eskdalemuir site, the calculated roughness
seriously disagreed with the subjective assessment, probably in this case because of
additional tree plantation since the time that the land use data was obtained. In these
cases, the subjectively assessed roughness length was used.
344 J. Ashcroft/J. Wind Eng. Ind. Aerodyn. 53 (1994) 331-355

At Aughton, Gravesend and Peterhead the anemometer height is greater than the
standard 10 m above ground because of the need to clear the disturbed wind flow
generated by high roughness elements in the vicinity (typically housing development).
At each of these sites the zero plane displacement was estimated as 0.7 of the
maximum height of buildings in the vicinity of the anemometer. The site diagrams
held in the Meteorological Office were consulted for this information. This value was
used to define the effective height for these sectors. If, at these sites, the sector
contained no contiguous built up area, the gust and mean wind speed were reduced to
10 m above ground using the power law profile with the appropriate power being
chosen from Ref. [6] according to terrain type. This adjustment resulted in a small
increase in the gust ratio for the sectors in question.
Where it was unclear whether a zero plane displacement should be used or whether
an adjustment of the gust and mean wind to 10 m should be undertaken (because of
the highly variable height of the roughness elements in the sector) then the gust ratios
for the sector were not taken into consideration in the overall analysis of the data. In
all cases the median gust ratio for any sector was not used if the sample size was too
small. A lower limit of 50 observations was thought a reasonable choice.

3.2.2. Results of the gust ratio-terrain roughness association


In order to present results compatible with the previous guidance on the terrain
dependence of gust ratios as given in Ref. [6], the median peak gust ratios were
plotted versus Kr, the terrain roughness parameter, which is defined as

in (105/Zor) In (10/Zo)
K~ = ln(105/Zo) ln(10/Zo~)" (10)

For the reference terrain where ZOr = 0.03 m and allowing for a general height with
zero plane displacement we have

Kr = 2.5855 ln[(z - d)/Zo]/ln(lOS/zo). (11)


(Note that Kr is the same as the parameter Se in Ref. [7] and most DALE stations
have d = 0 and z = 10m.)
The median gust ratios for the individual sectors at the 14 sites were plotted versus
the sector value of Kr and the best fit line drawn in by eye. The original plot of the
sector median values of the gust ratio G6o is reproduced here in Fig. 7. The scatter of
the data points is in part due to the fact that sector roughness value estimates
inevitably congregated around a limited number of standard values. In spite of this it
is still possible to define with some confidence the line of best fit for K r > 0.7. In the
highest roughness categories there are only a few gust ratio values available and the
line of best fit is more uncertain. The plotted data from this study agree well with the
data of Durst [2] and Deacon [3]. Standard estimates of z0 were used to describe the
terrain types for the data extracted from the above studies. The line of best fit suggests
a best estimate of the peak gust ratio G6o for terrain category 3 of 1.26, which is in
excellent agreement with that predicted by Vellozi and Cohen, taken from Ref. [6].
J. Ashcroft/J. Wind Eng. Ind. Aerodyn. 53 (1994) 331-355 345

G6x
1.5

1.4

X X X X X X
X X X X

X X XX X X XX X
XX X x ~(X X
1.3

, .xX
X X~ X l ~

1.2

1.1 Key: x Dale data


• Deacon (1965)
• Durst (1960)
• Velkazi and Cohen

1.0
0.5 0!6 0!7 o!e 0!9 1!0 1!1 1!2 11.3
Kr

Fig. 7. Sector median gust ratio G6o as a function of the terrain roughnessparameter Kr.

The lines of best fit for the gust ratios G600, G600c, G60 and Gp drawn to the DALE
data are reproduced in Fig. 8 along with the gust ratio-averaging time relationships as
given in Fig. A.1 of Part 1 of Ref. [6] or equivalently Fig. C.3.2a in Part 2 of Ref. [6].
The gust ratio G60 o can be seen to be only weakly dependent upon terrain
roughness. The DALE data suggests that the value of G6o o varies from 1.07 to 1.11
over the whole range of terrain roughness types found around the DALE stations. The
original data of Durst and Deacon was plotted on the original working figure and was
found to agree well with the DALE gust data (G6o o data only). Weiringa [5] derived
a mean value of G6o o of 1.10 for open country (corresponding to Kr = 1.0) but the
present results suggest a slightly lower value of 1.08 would be more appropriate. This
latter value is in good agreement with that of VeUozi and Cohen, appropriate for the
terrain type "open country", taken from Fig. C.3.2c in Part 2 of Ref. [6].
346 J. Ashcroft/J. Wind Eng. Ind. Aerodyn. 53 (1994) 331-355

The sector median values of the gust G6oo¢ , derived from a 10 min moving average
were similarly plotted up and the best fit line drawn through the scatter of data points.
This best fit line is reproduced in Fig. 8. The values of this gust ratio are some 2 to 4%
greater than the values of G6o o derived from six sequential 10 min averages. The gust
ratio G600c is however lower than that predicted by the relevant curve in Fig. C.3.2a of
Part 2 of Ref. [6], reproduced here in Fig. 8. It is not clear what data was used to
derive the above curve in Ref. [6]; it certainly lies above the best fit line drawn to the
gust ratios obtained in this study as well as those obtained in the previous studies of
Durst, Deacon, Vellozi and Cohen, and Weiringa.
The gust ratio derived from the DALE wind data for the 1 min average is also seen
to be consistently lower than the ratio suggested by Fig. C.3.2c in Part 2 of Ref. [6]
over the whole range of roughness categories. The reduction is typically 5%. It is
suggested that some previous estimates of the gust ratios may have been influenced by
the presence of unsteadiness in the mean wind and may have been overestimated, but
this would depend upon the mean wind speed. It is possible that previous estimates of
the gust ratios for winds of shorter duration than 60 s might be similarly have been
overestimated but unfortunately it is not possible to provide a definitive answer to this
question using the DALE data.
The peak recorded gust obtained from the Meteorological Office Mark 4 or Mark
5 anemometer system does not have a fixed averaging duration; the duration de-
creases with increasing wind speed. The line of best fit for the observed gust ratio Gp
is close to that of the 3 s duration gust ratio over most of the range of roughness.

2.2 Gust Ratio Key: BS 8100

2.1 Dale .....


peak , . ~ 600 MV: 10 minute moving average
2.0
gust "'%,%, ,, ,, ,, 600 6AV: 610 minute averages
1.9

1.8

1.7

1.6

1.5

1.4

1.3

1.2

1.1

1.0
o18 01~ o18 oI~ 11o 1~1 112 113
K~

Fig. 8. The gust ratio as a function of gust duration and the terrain roughness parameter K r. (Data from
BS8100 are reproduced with permission of BSI. Complete copies can be obtained from BSI sales, Linford
Wood, Milton Keynes M K I 4 6LE, UK.)
J. Ashcroft/J. Wind Eng. Ind. Aerodyn. 53 (1994) 331-355 347

(The position of the curve is rather less certain for Kr < 0.7.) It is not known whether
the original curve in Ref. [6-I was based on true 3 s average gust data or, as is more
likely, the peak recorded gust from the anemograph system was assumed to have a 3 s
duration. The peak gusts from the D A L E data (most hourly mean wind speeds in the
range from 11 to 20 knot) appear to be of a very similar, if perhaps slightly shorter
duration, but, according to the best fit line, greater than 1 s duration. Unfortunately,
there was insufficient data available by wind direction sector at wind speeds above 20
knot to determine the form of the relationship between G v and Kr at high wind speeds
when the effective averaging period of the cup anemometer system would be reduced.
The results of this study can be summarized in Table 1 which, derived using D A L E
data alone, gives the expected gust ratios for winds of 3, 60 and 600 s averaging period
(both overlapping and non-overlapping averages) as a function of terrain category at
standard height of 10 m above the zero plane. The terrain categories follow those
given in Ref. [6].

3.3. Dependence o f the ratio o f the peak gust to the 10 rain mean upon terrain roughness

It is of considerable interest to relate the peak recorded gust from the D A L E data to
the 10 min mean wind. This ratio has useful practical applications within the Me-
teorological Office as wind speeds reported from synoptic stations are averaged over
a 10 min period and it is also of interest to compare the ratios derived from D A L E
data with a number of predictive equations.
Firstly, it is necessary to determine whether there is any consistent difference
between the individual sector median values of the gust ratio Gp (non-centred 10 min
mean) and the median gust ratio G~,c (centred 10 min mean). The difference between
these two values was calculated for each station and sector. This was found to be very
small - the median gust ratio Gp was usually greater than Gp¢ by 0.01 or 0.02. The
overall mean difference over 14 stations was only 0.012 which is regarded as of no
great significance for practical applications. This result suggests that in practice, the
10 min mean does not have to be accurately centred around the time of the peak gust

Table 1
Gust ratios by averaging period and terrain category at 10 m height

Category 1 2 3 4 5
Zo (m) 0.003 0.01 0.03 0.10 0.30
Kr 1.21 1.11 1.00 0.86 0.72
Gp 1.44 1.49 1.56 1.66 1.85
G6o 1.21 1.23 1.26 1.31 1.37
G6o0 1.07 1.07 1.08 1.09 1.10
G6ooc 1.09 1.10 1.11 1.12 1.14

Explanation of terrain categories. 1: off-seawind onto flat coastal areas; 2: level grass plains, e.g. marsh; 3:
standard category: fairly level terrain-mostly open fields with a few houses and buildings; 4: fairly level
terrain with more hedges,trees and villages,farm buildings; 5: many trees and hedges,or fairlylevel wooded
country or more open suburban areas.
348 J. Ashcroft/J. Wind Eng. Ind. Aerodyn. 53 (1994) 331-355

for the median gust ratio G~ to be a good estimate (error + 0.01 compared to the
"true" estimate G~c). The very small difference between the two median values may be
explained by the fact that the peak gust of duration 2-3 s is almost invariably
occurring when the mean wind in that part of the hour is increasing and variations
from minute to minute in the wind are relatively small.
If the peak gust ratio (3 s gust) relative to the hourly mean wind is already known
for a particular location then it is possible to derive a first estimate of the peak gust
ratio relative to the 10 min mean wind by dividing Gp by the appropriate value of
G6o 0 for the terrain type, which may be taken from Table 1 or Fig. 8. These predicted
gust ratios were obtained by sector for each station and were correlated with the
t
actual median values of Gp. The correlation was excellent with r = 0.994, but with
a slight underestimate of the actual value. The suggested predictive equation is

Gpe = Gp/G6oo + 0.04. (12)


The very high correlation between Gp and Gpe is a reflection of the fact that the peak
gust is almost invariably associated with the maximum 10 min wind within the hour.
We choose here to present the plot of the gust ratio G'p but in this case the median
gust ratios are plotted against the parameter u' = ln[(z - d)/zo] in order to facilitate
a comparison with the predictive relation of Weiringa [5] and another derived from
that of Cook [6]. Weiringa [5] followed the method of Durst [2] and derived the
following equation to predict the value of the gust ratio Gp,
G~ = 1.0 + [1.42 + 0.3 I n ( T i t - 4)]/u'. (13)
Weiringa estimated the root mean square turbulent intensity of the wind to be 2.5u,,
where u, is the shear velocity, and assumed that the logarithmic wind profile was
valid, in which case the relative turbulent intensity, lu = 1.O/u'. The assumption that
I , = 1.O/u' appears to be roughly true if the whole spectrum of turbulent fluctuations
of periods from I to 3600 s is considered but it is not clear that this should apply to the
rms fluctuations over periods: t > few seconds and T < 3600 s, i.e. to a restricted part
of the spectrum.
To remove the possible uncertainty in assigning the duration t for short period
gusts as recorded by standard cup anemometer systems, Weiringa re-wrote T i t as the
number of gusts of wavelength l relative to the likely maximum gust wavelength
within the boundary layer, L ( ~ 1000 m). We then have
C~ = 1.0 + E/u', (14)

where the eccentricity, E = 1.42 + 0.3 In (lO00/l - 4).


For the U K Meteorological Office anemometer systems (Mark 4 and Mark 5)
Adams [13] estimated I as 84 to 97 m depending on wind speed. If we take I = 90 m we
then have
C~ = 1.0 + 2.0/u', or C~ = 1.0 + 2.01,. (15)
This equation is plotted in Fig. 9 along with the values of G~ as determined for each
sector with the value of u' based on the estimated value of Zo. The plotted data does
J. .4shcroft/J. Wind Eng. Ind. Aerodyn. 53 (1994) 331-355 349

B
2.2 i
Gp I.

2.1

2.0 Wll CI. ~

I
1.9
I' ~ x x
1.8 , x x~
~ ~ x
i \ x
1.7
~ t " x~
1.6-
~x~x.' x\ x
x
1 xx
%x
1.5 x , % ~ .
% xx
)oc
% x
% x x
1.4
% C
%
"~ ~. X ' W2
X xB
,%,
1.3

Key: Wl: Weidnga, L~1000 m, h,~90m


"" Wl
1.2 W2: Weiringa, T = 600, t = 3
C: Cook
B: Best fit to Dale data
1.1

I I I I I I I I
1,0 2.0 3.0 4.0 5.0 6.0 7.0 8.0
rj I

Fig. 9. The gust ratio G~,as a function of the terrain roughness parameter u'.

not agree very well with the peak gust ratios predicted by Eq. (15) - the data lies well
above the predictive line over the whole range of terrain roughness. Interestingly, if we
take Tit = 600/3 and use Eq. (13) directly to estimate the gust ratio the fit to the
plotted data is much improved over the whole range of terrain roughness and is
judged to be good for u' > 5.0. Unfortunately, the curve clearly lies above the body of
observed gust ratios over rougher terrain and is not a generally satisfactory fit overall.
The suggestion is that the estimates of the gust wavelengths l and L are somewhat
350 J. Ashcroft/J. Wind Eng. Ind. Aerodyn. 53 (1994) 331-355

erroneous as the introduction of these values into Eq. (13) leads to a substantial
underestimate of the gust ratio Gp.
Cook [7] adapted Weiringa's equation of the form of Eq. (13) by adjusting the
empirical coefficients so that the predicted gust ratio fitted a 1 s gust ratio of 1.6 over
open terrain (Zo = 0.03 m) at l0 m height. Cook assumed a peak recorded gust
averaging duration of 1 s. The dependence of the predicted gust ratio upon the terrain
class is implicit as the turbulence intensity is specified as a function of terrain
roughness. The equation is, in our notation,
Gp = 1.0 + 0.42Iu ln(3600/t). (16)
The turbulence intensity lu can be specified as a function of height in the boundary
layer and terrain roughness by the relation of Deaves and Harris as quoted in Ref. [7].
This equation has been considerably simplified here to
I, = 0.54/[F(u') u'] + 0.09, (17)
where
F(u') = 2.17 + 0.15u'.
The above equation is applicable only for z - d <<Zg and z - d >>Zo. If we substitute
(17) into (16) and divide Gp by the value of G6oo appropriate to the terrain type and in
addition use the small correction factor as given in Eq. (12) we can produce a predic-
tive equation for Gpt as a function of u'. We assume t = 3 s as an effective gust
averaging period for conditions when the hourly mean wind speed is from 11 to 20
knot, which covers the majority of occasions for which the gust ratio distributions
were derived.
The resulting equation is judged to be a good fit to the observed gusts G~ and lies
very close to the best fit line drawn in by eye for u' in the range from 4 to 7.5. Over very
smooth terrain the predicted gust ratio was slightly larger than the ratio suggested by
the best fit line and over rough terrain the DALE data suggest a rather higher value of
the gust ratio than predicted. The equation does however very satisfactorily describe
the form of the non-linear dependence of the gust ratio upon the roughness parameter
u' and is somewhat superior to relationships assuming an inverse dependence of
turbulent intensity upon the terrain roughness parameter u'. Note that the relation-
ship between the 3 s gust ratios and other longer period average gust ratios and
terrain roughness (Fig. 8) is also of a similar form.
The predicted 3 s gust ratio relative to the hourly mean as given by Eq. (16) was
derived and checked against the peak gust Gp taken from the best fit line to the DALE
data. This confirmed that Cook's predicted gust ratios (t = 3 s assumed) were very
close to the best estimate of Gp for terrain types 2 and 3 but were underestimates over
rough terrain (type 4 and especially type 5) and a small overestimate over smooth
terrain (type 1).
The values of G~ obtained from the line of best fit to the DALE data are shown in
Table 2.
An attempt was made to predict the values of G6oo and G6o as a function ofu' using
Eqs. (16) and (17) but the values obtained were far too large, almost certainly because
J. Ashcroft/J. Wind Eng. Ind. Aerodyn. 53 (1994) 331-355 351

Table 2
3 s gust to 10 min mean wind at 10 m height, by terrain category

Category 1 2 3 4 5
zo (m) 0.003 0.01 0.03 0.10 0.30
u' 8.11 6.91 5.81 4.61 3.51
Kr 1.21 1.11 1.00 0.86 0.72
G~, 1.36 1.42 1.48 1.58 1.74

the computed turbulent intensity lu represents the complete spectrum of wind fluctu-
ations and what is required is a partitioning of the turbulent intensity over the
relevant part of the spectrum (600 to 3600 s or 60 to 3600 s).

4. Dependence of the median gust ratio upon the hourly mean wind speed

Once the DALE data had been subject to stationarity checks it was possible to
determine whether any real dependence of the gust ratio upon the hourly mean wind
speed actually existed. Because of a generally insufficient sample size at high wind
speeds in the individual direction sectors it was not possible to study this phenomenon
as a function of terrain roughness (which is noticeably sector dependent). Instead, the
median gust ratios Gp, G6o and G6o 0 obtained by summing the distributions for all
eight sectors at each of the 14 study stations (wind speed class above 15 knot) were
related to those in the 11-15 knot class at each station. The trend line through the
arithmetic mean of the values in each wind speed class is plotted in Fig. 10. Around
the mean + / - one standard deviation is plotted.
Fig. 10a shows that there is a small and consistent decrease of the relative ratio of
G6oo and G6ooe as wind speed increases and this is very consistent across the whole
range of stations (small standard deviation). Note, however, that the gust ratio G6ooc
derived from the 10 min moving average data is rather more sensitive to the increase
in wind speed presumably because it much better reflects the variations in the shorter
period 1 min mean wind data used to form it. Because of the small variability of the
ratios within each class the small decrease in the mean ratio is statistically significant.
A t-test on the difference between the mean ratio of G6o o for the 16-20 knot class to
the 31-50 knot class revealed that the decline is significant at the 5 % probability level.
The magnitude of the reduction is rather small. A typical decrease of G6o 0 from
16-20 knot to > 30 knot is only 1%. The gust ratio at high wind speeds ( > 30 knot)
is only 1.5% less than that at 11-15 knot. This reduction is significant because the
absolute value of the gust ratio G6oo is not very sensitive to terrain roughness and so
there is little variation of G6oo within a wind class, i.e. between the majority of stations.
This decline is doubtless related to the small decrease in the relative turbulent
intensity as the mean wind speed increases 1-10].
The gust ratio at high wind speeds ( > 30 knot) is 1.5% less than the gust ratio
derived using the whole body of data (wind speed > 11 knot). This correction may be
used to adjust the roughness dependent gust ratios of averaging period 600 s, derived
352 J. Ashcroft/J. Wind Eng. Ind. Aerodyn. 53 (1994) 331-355

Ratio a)
1.00-

0.98

0.96

0.94
Key: • 10 min sequential
0.92
0 10 rain moving average
i i
6-10 11-15 16-20 21-25 25-30 31-50 Mean wind (kts)

Ratio b)
1.00-

0.98 -

0.96 -

0.94 -

0.92 -

0.90 -

I I
6-10 11-15 16-20 21-25 25-30 31-50 Mean wind (kts)

Ratio c)
1.02 -

1.00 -

0.98

0.96

0.94

0.92

I I
6-10 11-15 16-20 21-25 25-30 31-50 Mean wind (kts)

Fig. 10. The variation of the gust ratio with wind speed for: (a) 10 min mean wind; (b) 1 min mean wind;
(c) peak recorded gust.

for wind speeds > 11 k n o t to those appropriate to cases of hourly mean winds > 30
knot. The adjustment of the gust ratio G60oeto cases of high wind speeds is somewhat
greater. The gust ratio derived for all cases of hourly mean wind speeds > 11 knot
may be reduced by 2.0% to adjust it to that appropriate to conditions of very high
wind speeds, i.e. to the > 30 knot class.
The relative gust ratios G6o show rather greater variability within each speed class
as the shorter period gusts are more sensitive to the site roughness. Nonetheless there
is still a significant decline of G6o with increasing wind speed as the relative intensity of
J. Ashcroft/J. Wind Eng. lnd. Aerodyn. 53 (1994) 331-355 353

the wind fluctuations decreases. In this case the reduction of G6o is rather greater -
about 3% on average, as the mean wind speed increases from around 11 to 30 knot or
greater. The gust ratio G6o, derived from occasions of hourly mean wind speed > 11
knot for a chosen terrain type, may be reduced by 2.5°,/0 if it is required to estimate the
ratio appropriate to very high wind speeds ( > 30 knot).
The variation of the ratio Gp with respect to wind speed is rather more complicated.
The short period gust ratios are far more sensitive to local terrain roughness so there
is rather greater variability of the relative ratio within a wind speed class due to the
variability of roughness by sector at many stations and the differing terrain types
encountered at the 14 stations. In addition, it is known that the response time of the
anemometer decreases as the wind speed increases [1]. This may explain why the
relative gust ratios appear, on average, to increase after initially decreasing (this initial
decrease might be due to a reduction in the turbulent intensity with increasing wind
speed). The trend of the relative ratio at some stations contradicted the mean trend.
The relative gust ratio Gp at Benbecula, for example, tended to increase with increas-
ing wind speed. This might be thought to be due to the fact that the aerodynamic
roughness length of the sea surrounding much of the site increases as wind speed
increases but, conversely, the gust ratios for Sumburgh did not show any such trend.
It would be unwise to use the trend of the mean relative gust ratio as an indication
of the wind speed dependence of the 3 s gust ratio because of the large standard
deviation of the relative ratio due to the fact that individual sites can show their own
particular characteristics and of course the effective averaging duration of the peak
recorded gust is not constant at 3 s. Because of the inter-class variability in the relative
ratio the mean ratios by wind speed class do not show any statistically significant
decline from 16-20 to 26-30 knot nor is the increase in the mean relative ratio of
Gp from 26-30 to > 30 knot statistically significant.

5. Conclusions

The DALE 1 min wind data can provide very useful information on the dependence
of the gust ratio upon the terrain roughness for averaging periods from a few seconds
to 10 min and other interrelationships may be derived - for example the 3 s gust to
10 min mean ratio - but the study would have been more comprehensive if winds over
averaging periods from 10 to 30 s could have been recorded by the DALE stations. It
would have been then possible to provide a comprehensive assessment of the guidance
in Ref. [6] on the terrain dependence of gust ratios for all gust averaging periods from
a few seconds to 10 min.
The results in the present study suggest that the best summary guidance given in
Ref. [6] overestimates the 1 and 10 min gust ratios by roughly 5%, but the individual
gust ratios of these periods obtained from the data of Durst, Deacon, Vellozi and
Cohen and Weiringa are in fair agreement with the gust ratios derived from the DALE
data.
This work has obtained the empirical relationship between the gust ratio, averaging
duration and terrain roughness using a much larger body of data than has been
354 J. Ashcroft/J. Wind Eng. Ind. Aerodyn. 53 (1994) 331-355

available previously in any one study. The data covers a wide range of terrain
categories through the careful choice of stations. Though individual sector roughness
length estimates may well be erroneous, once sufficient data is accumulated the
general relationship between gust ratios and terrain roughness becomes clearer and
a best fit line can be drawn with some confidence for the majority of terrain types,
though users of the figures should note that less confidence can be placed in the trend
of gust ratios when the roughness length is > 40 cm. The gust ratios provided in this
study are derived in a consistent way and usefully update and augment the smaller
body of quoted information on the values of the gust ratio derived from various
sources as summarized, for example, in Ref. [4].
The peak gusts recorded at the Meteorological Office anemograph stations are
conservatively assumed to have an effective averaging duration of 3 s. A comparison
of the best fit line to the DALE peak recorded gusts with the best guidance in Ref. [6]
does not lead to a revision of this assumption. In addition, a comparison of the
observed peak gust to 10 min mean ratios with the gust ratio predicted from a rela-
tionship of Cook [7], assuming a gust duration of 3 s, gave good results over terrain of
moderate roughness. It is therefore suggested that the peak gust duration should not
be changed from its assumed value of 3 s as long as the hourly mean wind speed is
roughly 11 to 20 knot. It would be worthwhile to investigate, using anemograph data,
the relationship between the peak recorded gust and the hourly mean wind on
occasions of very strong winds, say 30 knot or more and to compare the gust
ratio-terrain roughness relationship with previous relationships given, for example,
for the 1 s gust.
It would be possible to further investigate the relationship of the short period gusts
to the hourly mean wind by defining the hourly mean wind as the average over an
hour centered around the gust time rather than the clock hour, which has been the
normal practice. This would, in practice, be quite difficult with the DALE archive as
there is no guarantee that, after quality control, the preceding and subsequent hours'
data will remain for processing along with the data for the current clock hour.
A longer period of DALE data ought to be processed to enable a sufficient body of
wind observations to be made available for such a study.
An important point brought out by the study is the need to undertake some check
on the temporal variation of the wind before calculation of the gust ratio. The DALE
1 min winds were most useful in this respect. It is believed that the stationarity checks
undertaken were most valid and very necessary. It may not always be possible to
undertake such checks in the routine analysis of anemograph data but, fortunately, it
appears that, when the hourly mean wind speed is greater than 20 knot, the gust ratio
may be considered to be effectively free from the effects of major unsteadiness in the
mean wind.

References

[1] Handbook of meteorological instruments, Vol 4. Wind instruments (Meteorological Office, HMSO,
London, 1980).
[2] C.S. Durst, Wind speeds over short periods of time, Meteorol. Mag. 89 (1960) 181-186.
J. Ashcroft/J. Wind Eng. Ind. Aerodyn. 53 (1994) 331-355 355

[3] E.L. Deacon, Wind gust speed: Averaging time relationship, Aust. Meteorol. Mag. 51 (1965) 11-14.
[4] C.E. Hardman, N.C. Helliwell and J.S. Hopkins, Extreme winds over the United Kingdom for periods
ending 1971, Climatological Memorandum no. 50A, Meteorological Office, Bracknell (1973).
[5] J. Weiringa, Gust factors over water and built-up country, Bound. Layer Meteorol. 3 (1973) 424-441.
[6] BS8100, Lattice towers and masts. Part 1: Code of practice for loading; Part 2: Guide to the
background and use of Part 1 (British Standards Institution, London, 1986).
[7] N.J. Cook, The designer's guide to wind loading of building structures, Part 1: Background, damage
survey, wind data and structural classification, Building Research Establishment, Garston (1985).
[8] R.R Brook and K.T. Spillane, The effect of averaging time and sample duration on estimation and
measurement of maximum wind gusts, J. Appl. Meteorol. 7 (1968) 567-574.
[9] M.E. Greenway, An analytical approach to wind velocity gust factors, Eng. Sci. Rep. 1241/78, Oxford
University Engineering Laboratory, Oxford (1978) (unpublished).
1-10] J. Ashcroft, Report on the study of gust ratios derived from DALE data and their dependence upon
averaging time, surface roughness and boundary layer stability, PSP Scientific Support Group
Technical Note no. 1, Meteorological Office, Bracknell (1992).
[11] J. Weiringa, An objective exposure correction method for average wind speeds measured at a shel-
tered location, Quart. J. R. Meteorol. Soc. 102 (1976) 241-253.
[12] M. Greengrass, Surface roughness information for the calculation of local wind climatologies,
Advisory Services Discussion Note no. 16, Meteorological Office, Bracknell (1989) (unpublished).
[ 13] R.J. Adams, Summary of theory of Weiringa's 'exposure correction factor', Met O. 3 Internal Report,
Meteorological Office, Bracknell (1984) (unpublished).

You might also like