Sie sind auf Seite 1von 36

Journal of Wind Engineering

and Industrial Aerodynamics 79 (1999) 233268


A two-site correlation model for wind speed, direction
and energy estimates
James R. Salmon*, John L. Walmsley
Zephyr North, 4034 Mainway, Burlington, Ont., Canada L7M 4B9
Atmospheric Environment Service, 4905 Duerin Street, Downsview, Ont., Canada M3H 5T4
Received 20 September 1997; accepted 2 April 1998
Abstract
A two-site wind correlation model was rst modied and then tested on long-term data from
ve pairs of Canadian weather stations. The stations were chosen to cover a variety of
topographic situations and surface cover types. A preliminary analysis of data from each of the
10 stations concluded that an absolute minimum of 12 months of monitoring is needed to
estimate the long-term wind speed distribution. Increasing this short-term monitoring period to
24 months causes noticeable improvements; thereafter, the improvement is more gradual. It was
also found that a short-term monitoring period between 12 and 24 months might produce
worse estimates than for a 12-month period. For the two-site tests, one station from each of the
ve pairs was designated as a reference station; the other was designated as a target station. It
was found that the model results derived from 1 yr of short-term simultaneous monitoring at
the two stations and long-term data at the reference station outperform the estimates based
solely on 2 yr of monitoring at the target station. The model results are further improved by
using 2 yr of short-termmonitoring. The conclusions from this study have implications for wind
energy and air pollution studies where long-term estimates of the wind speed and direction
distribution are needed and must be based on a relatively short-term period of monitor-
ing. 1999 Elsevier Science Ltd. All rights reserved.
Keywords: Two-site correlation; Wind energy estimates; Wind correlation model
1. Introduction
Walmsley and Bagg [1] developed a two-station wind correlation scheme and
applied it to data from Fort McMurray and Mildred Lake, Alberta. Simultaneous
hourly data from the period 197475 were used to produce a matrix of correlation
*Corresponding author. E-mail: zephyr.north@sympatico.ca.
Retired.
0167-6105/99/$ see front matter 1999 Elsevier Science Ltd. All rights reserved.
PII: S 0 1 6 7 - 6 1 0 5 ( 9 8 ) 0 0 1 1 9 - 6
coecients, which were multiplied by longer-term (196375) data from the Fort
McMurray Airport weather station to generate an estimate of the wind climatology at
Mildred Lake. (For a summary of the method, see Appendix A.) Walmsley and Bagg
[2] employed these synthetic data, comprising a joint relative frequency distribution
of wind speed, wind direction and atmospheric stability, in an application with the
climatological dispersion model (CDM) [3] to estimate surface concentrations of
sulphur dioxide.
The Walmsley-Bagg wind correlation model was developed to be compatible with
long-term, joint relative frequency data produced by the DayNight version of the
SAR program developed by the National Climatic Center at Asheville, North
Carolina. It was subsequently modied to accept a user-specied number of stability
classes and user-dened wind speed and direction classes, but until now it has not
been tested with independent data.
In addition to air pollution studies, the WalmsleyBagg model may also be applied
to the assessment of the potential long-termwind resource at a proposed wind-turbine
site. The wind-energy potential in the US Great Plains and Canadian Prairies is
enormous. In exploiting this potential, important questions that need to be answered
are: is a short-term dataset at a candidate turbine site representative of the long-term
climatology, and what is the best estimate of the long-term wind resource and
available wind energy.
The goal of the present study, therefore, was to test the WalmsleyBagg model on
independent data sets and to answer two specic questions: (i) What length of
short-term dataset at a candidate wind-turbine site is required to be representative of
the long-term climatology? and (ii) Can the model be used to improve the long-term
estimates of the wind resource at a candidate wind-turbine site?
2. Previous studies
A search of the literature failed to uncover any statistical correlation methods that
are closely related to the WalmsleyBagg method, although Walmsley [4] included
a short section on measurecorrelatepredict (MCP) methods. Of the methods
tabulated, only the WalmsleyBagg model had been published in the open literature.
The other [5] had been introduced in an internal report. Walmsley [4] also referred to
a formula given by AWEA [6] that may be used for estimating long-term mean wind
speeds at a potential wind turbine site. While an estimate of the mean may be useful in
a preliminary evaluation, information on the frequency distribution of wind speeds is
needed for a proper evaluation of the wind resource.
Kau et al. [7] introduced a statistical model for predicting surface wind speed and
direction. The former was found to depend primarily upon the slope wind, cross-
isobaric angle, surface thermal stability, and geostrophic wind; the latter depended
primarily upon the geostrophic wind direction, aspect angle of the topography,
up-canyon direction, and cross-isobaric angle in the boundary layer.
Hanna and Chang [8] derived an empirical formula (their Eq. (12)) for the spatial
correlation between wind speeds observed at two stations, as a function of spatial
234 J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268
separation and averaging time. It was found that their observations of spatial
correlations were best t if two independent integral distance scales were used:
a boundary-layer distance scale of about 300 m for small station separations and
a mesoscale distance scale of about 10 km for larger station separations. This formula
may be useful in determining whether a weather stations data applies at a nearby
prediction site. The formula, however, gives no estimate of the wind speed or direction
at the prediction site.
Palomino and Mart n [9] modied the interpolation scheme of Goodwin et al. [10]
by changing the weighting from an inverse-distance-squared formulation to the
dierence in topographic elevation between a grid node and an observation point.
This change tended to improve the spatial interpolation of wind elds in a complex
terrain. The method, however, was applied in a single valley and the maximum
horizontal distance between grid nodes and observation points was less than 2 km.
Finally, an abbreviated account of the present study appeared in Ref. [11].
3. Modications to the WalmsleyBagg model
For this study, the WalmsleyBagg model was modied slightly in order to
incorporate variable numbers and sizes of wind speed and wind direction classes or
bins. The number of stability classes (6) was unchanged and corresponds to the
standard distribution of DayNight SARstability categories. Thirty-two wind speed
classes with a bin size of 0.5 m s ranging from 0 to 16 m s were chosen. Sixteen
equally-sized wind direction bins of width 22.5 were selected, with the rst bin
centred on north. The total number of bins in the joint distribution was therefore
6;32;16, requiring a correlation matrix of size 3072` or 9.44;10", which was
successfully handled by our personal computer (PC) with 16 Mbytes of RAM. Later
tests showed a very weak dependence on atmospheric stability; therefore, it was
possible to eliminate stability while increasing the number of speed bins to 64 (i.e.,
ranging from 0 to 32 m s), making 64;16 bins and a correlation matrix of size
1024` or 1.05;10".
Wind speed bin widths of 0.5 m s correspond approximately to the size of bins
used for reporting and verication of wind-turbine power curves. The 16 wind
direction classes of width 22.5 correspond to the maximum allowed for use in the
MS-Micro wind ow model of Ref. [12]. They are also roughly the same size as the 12
classes of width 30 suggested for use in the WAP wind ow modelling program of
Ref. [13]. The software corresponding to the present model is designed so that the
user can specify the number (and hence the width) of speed bins, as well as the number
of direction bins, up to the limits discussed above.
4. Testing procedures
Five pairs of stations were examined. One station in each pair was designated as the
reference station and the other as the target station. In normal applications of the
J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268 235
WalmsleyBagg model, the reference station would be the one with long-term,
high-quality hourly wind data. (In this study, however, either station of the pair could
have been designated as the reference station, as all stations met the data require-
ments.) Atmospheric stability can usually be determined at the reference station from
other meteorological parameters using the DayNight SAR program. Air density
can be calculated from temperature and barometric pressure.
For these tests, a relatively short period of time is selected during which simulta-
neous hourly data are processed from both the target and reference stations. In
normal applications, wind data would be needed at the target station, temperature
and pressure may also be available, and parameters needed to calculate atmo-
spheric stability would likely not be available. Thus an assumption would be
made that atmospheric stability at the target station is the same as at the reference
station. If temperature is not measured at the target station, it would be estimated by
assuming a temperature lapse rate, which may be stability-dependent. If pressure is
not measured at the target station, it would be estimated from the reference station
conditions, using the hydrostatic approximation to adjust for any dierence in
elevation.
In normal applications, one would use the short-term data at the target and
reference stations together with the long-term data available at the reference station
and apply the WalmsleyBagg model to estimate the long-term wind climatology as
a function of stability at the target station. (By incorporating a calculation of air
density, an estimate of the long-term available wind power can also be determined.) In
the present case, because of the availability of long-term data at both stations, the
process was taken one step further: verication of the estimated wind climatology and
wind power with long-term data at the target station.
In summary, in the present study, we applied the modied WalmsleyBagg model
to pairs of stations for which the long-term data sets were available at both stations
and designated one station as the reference and the other as the target. The estimated
long-term wind climatology at the target was then compared with the observed
climatology.
5. Stations and periods for testing
Five pairs of stations were chosen for testing. They are shown in Fig. 1 and listed in
Table 1. (Hereafter the A in each station name, used to identify the airport site, will
be dropped for brevity, as will the International following Vancouver.) The
criteria for choosing the station pairs were: proximity, geographical distribution,
variety of terrain and surface cover, and data quality and stability. The focus was
restricted to Canada because quality-controlled data were readily available, but the
decision to consider a variety of conditions implies that the results will have wider
applicability.
Fig. 1 and Table 1 show that the stations in each pair are reasonably close to each
other (within 100 km, with one exception), increasing the likelihood that both the
reference and target stations were within the same air mass and therefore had similar
236 J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268
Fig. 1. Map of Canada showing the ve pairs of weather stations. Inset maps are, respectively, from east to
west, Regions AE of Table 1.
Table 1
List of stations with period, location, elevation and distance between each pair. The rst station listed in
each region was used as the reference station; the second is the target station. All data periods ended on 31
December 1989
Region Description Station Start of
Period
Lat.
(N)
Long.
(W)
Elev.
(m MSL)
Dist.
(km)
A Northumberland Moncton A 1968 46.12 64.68 71
Strait Summerside A 1968 46.43 63.83 24 74
B St. Lawrence Mont Joli A 1969 48.60 68.20 52
Estuary Baie Comeau A 1967 49.13 68.20 22 59
C Great Lakes Wiarton A 1972 44.75 81.10 222
Gore Bay A 1972 45.88 82.57 193 170
D Prairies Moose Jaw A 1965 50.33 105.55 577
Regina A 1969 50.43 104.67 577 64
E Lower Fraser Vancouver Int. A 1964 49.18 123.17 3
Valley Abbotsford A 1972 49.03 122.37 58 61
J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268 237
atmospheric stability characteristics and that their wind regimes were probably
correlated. It is evident that the ve station pairs represent ve well-dened geo-
graphical regions (designated AE) of Canada. (The middle and far north, however,
lack representation.) The terrain types range from maritime and at (Region A) to
continental and rolling (Regions C and D) and from prairie grassland (Region D) to
lakeside forest (Region C). The wind ow may be strongly inuenced by the presence
of larger-scale topography: mountains along either side of the estuary in Region B and
the river delta bordered by mountains on the north and southeast sides in Region E.
By considering this variety of conditions, it is expected that the results will have
applicability elsewhere, both within Canada and in other countries.
Regarding data quality, an attempt was made to choose stations that had no
obvious data aws. Treatment of the data to eliminate potential problems is described
below.
6. Input data processing
Data sets for this study were available in the Canadian Weather Energy and
Engineering Data Sets (CWEEDS) compiled by the Atmospheric Environment Ser-
vice (AES) in 1993. This large compilation of Canadian weather data at 143 stations is
available on CD-ROM in compressed WYEC2 (Weather Year for Energy Calcu-
lation, Version 2) format. A companion diskette to the CD-ROM provides anemom-
eter-height history data.
The data required for the present study were wind speed, wind direction and the
parameters required to calculate atmospheric stability (see below). Temperature and
station barometric pressure were also extracted from the CWEEDS data sets as
auxiliary data.
6.1. Wind speed and wind direction
Wind speed and direction data were extracted directly from the CWEEDS WYEC2
les. Both required further processing, however, to minimize the discretization prob-
lems associated with the use of integer values in recording, converting and storing the
data.
In past years, wind speed measurements were recorded as values (in either
miles h, kt or, more recently, in km h) to the nearest integer. A conversion was
often made, e.g., from miles h to kt and the integer nearest to the converted value
was stored. All processed data in the AES archive have since been converted to integer
values in units of km h. For the CWEEDS data sets, they were further converted to
m s and stored as integer values in dm s. These recording, conversion and
storage procedures could potentially have caused problems for data analysis, as gaps
appeared in the continuum of wind speeds. In converting between integer values from
kt to km h, for example, the only possible resultant wind speeds are 0, 2, 4, 6, 7, 9,
11, 13, 15, 17,
2
km h. A problem arises, therefore, when the distribution of wind
speeds is examined. If, for example, the data are placed in wind speed bins of width
238 J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268
1 km h, there would be no data at all in bins representing 1, 3, 5, 8, 10, 12, 14, 16,
2
km h, and the representation of the distribution of the wind speeds would be
unrealistic.
To avoid the problem of empty wind-speed bins, we treated the data as follows.
Each wind speed datum was converted back from its CWEEDS value in dm s to its
AES archive value in km h and thence to an integer value in the units (miles h or
kt) originally used for recording. Since the corresponding forward conversion proced-
ures went from lower to higher resolution, thus creating empty bins, it was always
possible to recover the originally recorded integer from this reversal process. To this
original value was added a random oset so that the modied values ranged over the
appropriate integer-bin width. The modied number was then subjected to the
original conversion process but without rounding to the nearest integer values at any
stage. The result of this procedure was a continuum of wind speeds in the archived
units.
The wind direction data were subjected to a similar procedure. Each value was
converted from its archive storage value (i.e., to the nearest 10 value) back to its value
in the original measurement system, either 8 or 16 sectors. This was done using a
conversion table supplied by AES: Documentation for the Digital Archive of Canadian
Climatological Data (Surface) Identied by Element. The resulting original data had
random osets introduced to spread them evenly within their (8 or 16) sectors. These
converted data were then used for the analysis.
Temperature and barometric pressure data were copied from the CWEEDS data
sets unchanged.
6.2. Atmospheric stability
As atmospheric stability data are not recorded explicitly in the CWEEDS data sets,
they were calculated using a set of AES customized programs modied to run on PCs.
First, the annual data les were extracted from the CD-ROM and uncompressed
(using PKUNZIP) to the PC hard drive. The les for the period under study (which
varies by station; see Table 1) were then concatenated into a large ASCII le in WYEC2
format. This le was subsequently converted to a proprietary format readable by the
AES customized programGRP115S as modied for use on a PC. GRP115S calculates
hourly atmospheric stability from station location, date and time, cloud ceiling height,
wind direction, wind speed and amount of cloud cover. All of these, except for the
location data, are included in the CWEEDS data sets. GRP115S uses the SAR
algorithm to determine which one of six stability classes should be assigned to the
data hour in question. The six stability classes were introduced by Pasquill [14] and
are known as the PasquillGiord stability categories.
6.3. Further data processing
The two data sets, one with the atmospheric stabilities and the other with the wind
speed, wind direction, temperature and barometric pressure were merged into one
data set of hourly values of these parameters.
J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268 239
Fig. 2. Autocorrelation results obtained by using short-term data for varying lengths of time at a single
station to estimate the long-term climatology at the same location. (a) Worst-case estimates, X
'`
/
''
and
N
'`
/
''
. (b) Standard deviations of estimates, 1$
'`
/
''
.
The data were then checked for continuity (no signicant gaps) and completeness
(24 h coverage). Also checked were the anemometer histories of the stations to ensure
that the anemometers had not been moved (horizontally or vertically) during the data
periods of the study.
The above procedures resulted in acceptably clean data sets for the statistical
studies to be described below.
7. Autocorrelation results
The rst set of tests was carried out on each of the 10 station data sets in Table 1,
i.e., there was no segregation into reference and target locations. The purpose was to
240 J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268
Fig. 2. Continued.
determine the sensitivity of the long-term average wind speed estimate based on the
length of the short-term observation period. These results were then used to judge the
accuracy of the correlation method.
The following parameters were dened:
''
and
'`
are the average wind
speeds over the long- and short-term data records, respectively, at the target station;

'`
, N
'`
and X
'`
are the standard deviation, minimum and maximum, respectively,
of
'`
based on a series of calculations over the short-term record. In these auto-
correlation tests,
''
was based on at least 18 and up to 26 yr of data. For
'`
,
the period ranged from 1 to 60 months. Since there are a number of contiguous
short-term periods within a long-term period (e.g., 216 one-month periods in an 18 yr
record),
'`
was calculated for each of these. (Note that all possible contiguous
J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268 241
short-term periods were used, resulting in an overlap in the case of short-term periods
of greater than one month in length. Thus, for example, there were 205 contiguous
12-month periods in 18 yr.) The outcome of this test was a series of estimates of

''
based on values of
'`
determined from short-term data periods of varying
length.
The results of the autocorrelation calculations are presented in Fig. 2. Fig. 2a
shows the ratios of X
'`
/
''
and N
'`
/
''
and Fig. 2b shows values of 1$
'`
/
''
, all
as functions of the number of months in the short-term period. All values are
presented as percentages.
Fig. 2a provides worst-case estimates of the long-term wind speed as a function of
the length of the short-term monitoring period. As would be expected, the worst-case
estimates improve, i.e., approach 100%, as the short-term period increases. The
improvement is dramatic up to 1 yr and then gradual thereafter. The worst worst-case
estimates at one month are 35% (Abbotsford) and 180% (Wiarton). These improve to
76% (Baie Comeau) and 126% (Abbotsford) after 1 yr and to 84% and 120% (Mont
Joli) after 2 yr. The best worst-case estimates at one month are 65% and 133%
(Moncton), which improve to 87% and 114% (Moncton) after 1 yr and to 91% (Gore
Bay) and 107% (Wiarton, Moncton) after 2 yr. In all cases there is a gradual, but
steady improvement in the worst-case estimates after the second year of short-term
monitoring.
Fig. 2b shows the behaviour of the standard deviation of the estimates. The plots
showa behaviour similar to the worst-case curves of Fig. 2a. Values of
'`
/
''
ranged
between 13% (Regina) and 26% (Abbotsford) at one month, between 5% (Moncton)
and 13% (Abbotsford) at 1 yr and between 3% (Wiarton) and 12% (Abbotsford) at
2 yr. For comparison purposes, Table 2 shows percentage uncertainties in estimates of
long-term mean wind speeds from other studies. At the 95% condence interval, these
are 1113% after 1 yr of monitoring and 89% after 2 yr, i.e., in the upper half of the
ranges shown in Fig. 2.
The curves for Moncton are reproduced in Fig. 3 at higher resolution as an
example to show that the estimates are worse in the 1322 month monitoring period
than they are at 12 months. This is attributed to seasonal variations and the use of
a non-integral number of years to estimate the long-term annual average from the
short-term calculations.
The results of the autocorrelation tests suggest the following procedures for
estimating the long-term wind regime at a given site. A minimum of 12 months of
monitoring is recommended. Even with 12 months of monitoring, there is still
signicant uncertainty in the estimates of the long-term average speed (
'`
/
''
ranging from 513%; worst errors ranging from 1326%). If these uncertainties
are considered unacceptable, 24 months of monitoring would reduce the stan-
dard deviations slightly (
'`
/
''
to the 412% range) and the worst errors more
signicantly (to 720%). After 24 months, the improvement is more gradual. A
decision to monitor for a period between 12 and 24 months could produce
worse estimates than for the 12-month period. The same is true for the 2436,
3648 and 4860 month periods, though this eect of seasonal variability diminishes
with increasing length of record. Nevertheless, it is important to monitor for an
242 J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268
Table 2
Percentage uncertainties in estimates of long-term mean wind speed as
a function of monitoring period (source: Ref. [4])
Condence Source Period (months)
interval (%)
1 12 24 60
80 WR83 24 7 5 3
90 C80 37 11 8 5
95 WR83 37 11 8 5
95 C80 44 13 9 6
WR83 from Ref. [16].
C80 from Ref. [17]. See also Ref. [18].
Fig. 3. Autocorrelation results for Moncton. Worst-case and standard deviations of estimates, plotted at
higher resolution than the curves labelled T in Fig. 2.
J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268 243
Fig. 4. Cross-correlation results obtained by using simultaneous short-term data for varying lengths of
time at both the reference and target stations and long-term data at the reference station to estimate the
long-termclimatology at the target station using the modied WalmsleyBagg model. Plotted are values of
1$
'`
/
''
. The station names at the top of each diagram indicate the reference and target locations,
respectively. (a) Northumberland Strait. (b) St. Lawrence Estuary. (c) Great Lakes. (d) Prairies. (e) Lower
Fraser Valley.
244 J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268
Fig. 4. Continued.
integral number of years to get the best estimate of thelong-term wind climatology.
These results also suggest that at least 12 months, and preferably 24 months of
short-termdata are required for the two-site correlation application to be described in
the next section.
8. Two-site correlation results
8.1. Standard deviation of estimates
The wind-speed results from application of the modied WalmsleyBagg model to
each of the two-site reference-target pairs are presented in Fig. 4. Only the standard
deviations of estimates are shown, as values of 1$
'`
/
''
, corresponding to the
autocorrelation results of Fig. 2b, which displays similar looking curves.
Use of the modied WalmsleyBagg correlation model (solid curves in Fig. 4)
reduced the standard deviations compared with what would be obtained by
monitoring alone (dashed curves), i.e., inferring wind climatology from short-term
monitoring without the use of a model. The improvement is signicant for monitoring
periods of less than 1 yr in all regions except the Prairies (Fig. 4d), where it is just
noticeable.
Standard deviations after 12 months or more of monitoring without model ap-
plication (dashed curves) are in the range 410% in Fig. 4a4d and 1013% in
Fig. 4e. For monitoring periods of 12 months or more, application of the model
J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268 245
caused reductions of 01 percentage points in Fig. 4a and 4d, 1 percentage point in
Fig. 4c and 23 percentage points in Fig. 4b and 4e. These translate to reductions in
standard deviation of 020% in Fig. 4a, 2035% in Fig. 4b and 1020% in Fig. 4c
and 4e. Even though the model-enhanced results in Fig. 4e are still consistently poorer
than those of Fig. 4a4d, they are signicantly improved over the monitoring-only
scenario. The correlation model gave little improvement over 12 months or more of
monitoring in Fig. 4d.
As noted, even after application of the correlation model, the results in Fig. 4e are
consistently poorer than those of the other four reference-target pairs. This suggests
that wind data from the Abbotsford and Vancouver stations are the most poorly
correlated of the ve pairs of station data selected for this study.
Fig. 5 presents, as an example, both the worst-case and the standard devia-
tions of estimates for the pair of stations in the Northumberland Strait area (see
Fig. 4a). As with the standard deviations (Fig. 4), the worst-case estimates were
improved most signicantly by application of the model in the rst 12 months of
monitoring.
There was very little dierence amongst the three methods of estimating long-term
average wind speed: using the model to correlate (i) wind speed alone; (ii) wind speed
and direction; and (iii) wind speed, wind direction and thermal stability. This is shown
by solid curves in Fig. 4 for the rst two of these methods. Not shown but equally true
are the results of correlating speed, direction and stability.
8.2. Correlation matrices
Tables 35 present sample long-term correlation matrices for stability, wind
speed and wind direction, respectively, using Moose Jaw and Regina as the reference-
target pair. For display purposes, the 32 wind-speed bins of width 0.5 m s were
reduced to ve of width 4 m s and the number of wind direction classes was reduced
from 16 to 8. The results presented here, however, are very similar to those produced
with greater resolution. Summing the values in the cells along the central diagonal
gives the percentage of time that the same class occurred at both stations. The
footnote at the bottom of each table indicates that the same class was observed
4559% of the time. The frequency of classes that diered by no more than one was
obtained by summing values along the three central diagonals. (Recalling that the
wind direction is a circular distribution, Class 8 is adjacent to Class 1, so the top-right
and bottom-left cells are included in the summation.) These frequencies ranged from
8098%.
Table 6 summarizes the two-site correlation matrices and presents additional
statistics for the individual regions. Table 6a shows that the same stability class occurs
at the reference and target stations 5070% of the time. The same or next adjacent
class occurs 8095% of the time. The highest values are found in Region A; the lowest
in Region D. The dierences found in Region D, the least geographically complex of
the ve regions, were not expected.
Table 6b shows that the same wind speed class occurs at the reference and target
stations 5268% of the time. The same or next adjacent class occurs 9598% of the
246 J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268
Fig. 5. Cross-correlation results for Northumberland Strait (see Fig. 4a). The six inner curves show values
of 1$
'`
/
''
; the six outer curves show worst-case estimates, X
'`
/
''
(upper three curves) and N
'`
/
''
(lower three curves).
time. Signicant dierences in these frequencies between regions are not apparent.
Also shown are mean wind speeds at both reference and target stations for the
short-term (2 yr) and long-term (full dataset) periods and the correlation coecient
between reference and target datasets for both short- and long-term periods. The
mean wind speeds range between 2.6 and 5.5 m s, with the highest values tending to
be in Regions A and D (and at the reference station in Region B) and the lowest in
Region E. The correlation coecients are highest (&70%) in Regions A and D and
lowest (&35%) in Region E.
Table 6c shows that the same wind direction sector occurs at the reference and
target stations 1649% of the time. (With eight sectors and random distributions, the
J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268 247
Table 3
Sample reference-target thermal stability correlation matrix for Region D. The period is 196990. Six
PasquillGiord stability categories, ranging from unstable (1) to stable (6), are shown. The reference
station (columns) is Moose Jaw; the target station (rows) is Regina. N"182 990. Row, column and
diagonal totals are subject to round-o errors
Class 1 2 3 4 5 6 Total
1 0.000 0.001 0.001 0.002 0.000 0.000 0.003
2 0.001 0.006 0.008 0.024 0.000 0.000 0.039
3 0.001 0.010 0.016 0.056 0.002 0.002 0.087
4 0.002 0.024 0.049 0.437 0.052 0.055 0.621
5 0.000 0.000 0.002 0.087 0.020 0.021 0.131
6 0.000 0.000 0.002 0.079 0.019 0.020 0.120
Total 0.004 0.041 0.078 0.685 0.093 0.098 1.000
Sum on diagonal"0.499; sum on three central diagonals"0.805.
Table 4
Sample reference-target wind speed correlation matrix for Region D. The period is 196990. Five wind
speed classes with bin widths of 4 m s are shown. The reference station (columns) is Moose Jaw; the target
station (rows) is Regina. N"182 990. Row, column and diagonal totals are subject to round-o errors
Class 1 2 3 4 5 Total
1 0.266 0.108 0.009 0.000 0.000 0.384
2 0.135 0.250 0.047 0.003 0.000 0.435
3 0.009 0.070 0.068 0.008 0.000 0.155
4 0.000 0.003 0.012 0.008 0.001 0.024
5 0.000 0.000 0.000 0.001 0.001 0.003
Total 0.410 0.431 0.136 0.020 0.003 1.000
Sum on diagonal"0.593; sum on three central diagonals"0.975.
Table 5
Sample reference-target wind direction correlation matrix for Region D. The period is 196990. Eight wind
direction classes with bin widths of 45 are shown. Class 1 is centered on N, Class 2 on NE, etc. The
reference station (columns) is Moose Jaw; the target station (rows) is Regina. N"182 990. Row, column
and diagonal totals are subject to round-o errors
Class 1 2 3 4 5 6 7 8 Total
1 0.032 0.010 0.003 0.002 0.002 0.003 0.010 0.027 0.089
2 0.010 0.015 0.008 0.003 0.002 0.002 0.003 0.005 0.048
3 0.004 0.011 0.039 0.024 0.007 0.006 0.005 0.005 0.100
4 0.004 0.004 0.036 0.101 0.051 0.030 0.014 0.007 0.247
5 0.001 0.001 0.003 0.012 0.028 0.025 0.009 0.003 0.084
6 0.001 0.001 0.001 0.002 0.009 0.036 0.021 0.004 0.076
7 0.003 0.001 0.002 0.002 0.004 0.026 0.095 0.029 0.161
8 0.018 0.003 0.002 0.002 0.002 0.006 0.059 0.102 0.196
Total 0.073 0.047 0.094 0.148 0.105 0.134 0.217 0.182 1.000
Sum on diagonal"0.448; sum on three central diagonals#two corner cells"0.823.
248 J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268
Table 6
a. Summary of thermal stability correlation in Regions AE
Variable Source Region
A B C D E

Table 3 0.707 0.649 0.635 0.499 0.603

`
Table 3 0.945 0.901 0.888 0.805 0.851

"Sum on central diagonal.

`
"Sum on three central diagonals.
b. Summary of wind speed (m s) correlation in Regions AE

Table 4 0.589 0.521 0.594 0.593 0.676

`
Table 4 0.977 0.954 0.976 0.975 0.976

"`
Fig. 7a 4.5 5.1 4.1 5.3 3.8

'`
Fig. 7a 5.2 4.1 4.9 5.5 2.8

"'
Fig. 7b, Fig. 9 4.7 5.2 3.9 5.0 3.2

''
Fig. 7b, Fig. 9 5.5 4.4 4.6 5.3 2.6
r
`
0.73 0.55 0.56 0.69 0.33
r
'
0.72 0.55 0.56 0.68 0.37

"Sum on central diagonal.

`
"Sum on three central diagonals.
"Mean wind speed.
R"Reference; T"target.
S"Short-term (2 yr); L"long-term.
r"Correlation coecient, R and observations.
c. Summary of wind direction () correlation in Regions AE

Table 5 0.492 0.320 0.357 0.448 0.161

`
Table 5 0.879 0.708 0.737 0.823 0.498

"`
$
"`
248$92 224$96 239$112 271$105 127$99

'`
$
'`
241$102 286$99 269$120 264$137 189$143

"'
$
"'
241$88 235$92 239$103 253$98 123$99

''
$
''
240$95 289$97 265$110 231$129 218$122

'`"`
$ !3$36 12$64 !3$52 !4$42 !24$87

''"'
$ 4$35 9$62 1$52 !2$43 !15$89

"Sum on central diagonal.

`
"Sum on three central diagonals#corner cells.
"Circular mean wind direction.
"Circular standard deviation of wind direction.
R"Reference; T"target; S"short term (2 yr); L"long term.
TSRS, TLRL"statistics based on wind direction dierences.
J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268 249
Fig. 6. Sample wind speed and direction distributions for the Northumberland Strait region. Model
estimates use wind speed and direction correlation. (a) 1 yr period of simultaneous hourly data at the
reference (Moncton) and target (Summerside) stations. Dierences between target and reference distribu-
tions appear in the bottom pair of plots. (b) 22 yr period of data at the reference site and a comparison of
estimates and data for the same long-term period at the target site. Dierences between estimates and data
distributions at the target site appear in the bottom pair of plots.
sum along the central diagonal would be 12.5%.) The same or next adjacent
sector occurs 5088% of the time (compared with 37.5% for random distributions).
The highest values are found in Region A; the lowest in Region E. Values of the
circular mean wind direction and their respective circular standard deviations [15]
250 J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268
Fig. 6. Continued.
J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268 251
(see Appendix B) are also presented in Table 6c for the reference and target
stations over both short- and long-term periods. Circular means and standard
deviations were computed for the dierence in wind directions (target reference) for
both periods and appear in Table 6c. The nonlinear nature of these data is evident, as
the dierences in mean wind direction are not the same as the means of the dierences.
Mean direction dierences are )12 in Regions AD but are signicantly larger
(&20) in Region E.
9. Estimated wind speed and direction distributions
9.1. Sample distributions. Region A: 1 yr short-term period
Fig. 6 presents another method of displaying the results, using Region A as an
example. The top pair of plots in Fig. 6a shows the wind direction and wind speed
distributions at the reference station (Moncton), based on hourly data for a 1 yr
period (1968). The middle pair of plots displays the target station (Summerside) data
for the same period. The bottom pair gives the dierences between the target and the
reference station data. Wind direction and speed bins are 22.5 and 0.5 m s wide,
respectively. The low frequency in the second speed bin (0.51.0 m s) compared with
the rst and third bins is possibly related to the anemometer start-up speed, rather
than to a physical phenomenon. This gap would be signicantly reduced if bins of
width 1.0 m s were used. There is also a spike in the 6.57.0 m s bin in the
Moncton data. This persists out to 8 yr of observations (196875) and is probably
related to observer bias, 7 m s being close to both 15 miles h and 25 km h.
This bias probably does not have much impact on the present calculations, as nothing
signicant shows up in the target estimate minus actual plots in Fig. 6b.
The top pair of plots in Fig. 6b displays the long-term (22 yr) data at the reference
station (Moncton). The wind rose is somewhat dierent from the corresponding
short-term period (Fig. 6a), the maximum frequency at W having been shifted to
WSW. In the wind speed distribution, the main dierence from the short-term is
a broadening and lowering of the peak frequencies and elimination of the spike near
7 m s. The third pair of plots in Fig. 6b shows the observed long-term data at the
target station (Summerside). Here, the wind speed distribution is very similar to the
short-term, except for a drop in frequency of the lowest speed bin from 104%. The
short-term and long-term wind roses at Summerside look rather similar, except for
a decrease in frequency of westerly winds from 149%.
The wind rose at Moncton is signicantly dierent from that at Summerside, both
for the short- and the long-term periods (Fig. 6a and 6b, respectively). Monitoring the
wind direction at one station would not give a good estimate of the wind rose at the
other station without use of the wind correlation model.
Examining the plots in the second row from the top of Fig. 6b, it can be seen that
the model-produced target station estimate of both wind speed and direction
distributions more closely resemble the corresponding target station actual than
they do the reference station, i.e., the wind climatology at Summerside cannot be
252 J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268
inferred from wind monitoring at Moncton. In fact, monitoring at Moncton for 22 yr
without the application of the model gives a poorer estimate of the long-term wind
climatology at Summerside than does 1 yr of simultaneous data with the wind
correlation model.
The dierences between estimated and observed distributions at Summerside are
displayed in the bottom pair of plots in Fig. 6b. Dierences in wind direction
frequencies do not exceed 2.5 percentage points for any of the 16 sectors. Dierences
in wind speed frequencies are less than 1.4 percentage points in magnitude, except for
the lowest wind speed bin where the dierence is 5 percentage points.
Fig. 6a and 6b also give observed and estimated values of the average wind
speed and the kinetic power density, dened in the same way as wind power
density but with air density held xed at 1 kg m`. At Moncton, the short-term
values of these parameters agree quite closely with the long-term values (4.7 m s
and 103 W m`, respectively). At Summerside, however, both the short-term data
and the long-term model estimate underpredict the observed long-term average
speed (5.5 m s) and kinetic energy density (173 W m`) by 57% and 813%,
respectively.
9.2. Sample distributions. Region A: 2 yr short-term period
Fig. 7 portrays the same information as Fig. 6, except that the short-term period is
two years instead of one. The main dierences found in the 2 yr data (Fig. 7a)
compared with the 1 yr data (Fig. 6a) are a reduction in the frequency of westerly
winds at both stations and a reduction in frequency of the lowest wind speed bin at
Summerside. The target station estimate wind rose in Fig. 7b is very similar to the
corresponding one in Fig. 6b. The most obvious improvement in the wind speed
distribution is the reduction of frequency in the lowest bin, cutting the deviation from
the observed frequency from 5.2 to 2.6 percentage points. The underprediction of wind
speed is reduced from 5% to 2% and the underprediction of kinetic energy density is
reduced from 13% to 6%.
9.3. Sample distributions. Region A: 4 yr short-term period
Fig. 8 is the same as Figs. 6 and 7, except that the short-term period is now 4 yr.
The wind rose in Fig. 8a remains much the same as it was in Fig. 7a, while the
wind speed distribution shows a further small reduction in frequency of the
lowest wind speed bin at Summerside. The target station estimate wind rose in
Fig. 8b is somewhat better than the corresponding ones in Fig. 6b and Fig. 7b,
especially in the frequency of southerly winds. There is a further improvement
in the lowest wind-speed bin, the deviation from the observed frequency being reduced
to 1.6 percentage points. The underprediction of wind speed remains at the same level
as in Fig. 7b (2%), whereas the underprediction of kinetic energy density is further
reduced to 5%. Increasing the short-term period of simultaneous observations be-
yond 4 yr did not yield signicant improvement over the 22 yr estimates shown in
Fig. 8b.
J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268 253
Fig. 7. Same as Fig. 6 except that the short-term period is 2 yr. (a) Short-term data and dierences between
target and reference distributions. (b) Long-term data, estimates and dierences.
9.4. Distributions for Regions BE: 2 yr short-term period
Fig. 9 shows long-term wind speed and direction distributions for the remaining
four regions. The short-term period is 2 yr, so these gures should be compared with
Fig. 7b. Examining Fig. 7b and Fig. 9 reveals several interesting facts. First, the
observed wind roses at the reference and target stations are very dierent from one
another. Even in the Prairies (Fig. 9c), where the closest resemblance is found, the
254 J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268
Fig. 7. Continued.
J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268 255
Fig. 8. Same as Fig. 7 except that the short-term period is 4 yr. (a) Short-term data and dierences between
target and reference distributions. (b) Long-term data, estimates and dierences.
frequency of SE winds at Regina is double what it is at Moose Jaw. It is clear that
monitoring wind direction at the reference site, even for a long period of time, would
not give a good estimate of the wind rose at the target site and vice versa. Second,
except in the Prairies, the observed wind speed distributions at the reference and
target sites are signicantly dierent. This is also reected in the values of average
wind speed and kinetic power density. Again, long-term monitoring of wind speed at
the reference site would give a poor estimate of the wind speed distribution, average
256 J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268
Fig. 8. Continued.
J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268 257
Fig. 9. Wind speed and direction distributions for all regions except Northumberland Strait. Short-term
period is 2 yr. Long-term data, estimates and dierences are displayed. Model estimates use wind speed and
direction correlation. (a) St. Lawrence Estuary. (b) Great Lakes. (c) Prairies. (d) Lower Fraser Valley.
258 J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268
Fig. 9. Continued.
J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268 259
Fig. 9. Continued.
260 J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268
Fig. 9. Continued.
J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268 261
wind speed and kinetic power density at the target site and vice versa. Third, use of
the wind correlation model gives a much better estimate of the long-term wind
climatology at the target station, even with only 2 yr of short-term data, compared
with monitoring at the reference station alone. Dierences between estimated and
observed frequencies at the target site were within 3.9 percentage points for all
direction sectors and speed bins; in the great majority of cases, they were within 1.8
percentage points.
9.5. Summary
Table 7 summarizes the results of estimating long-term average wind speed and
kinetic power density at the ve target stations based on model calculations that
used 1, 2 and 4 yr of short-term data. Table 7a gives the values, while Table 7b
presents the percentage dierences from the long-term observed values. Several facts
emerge from an examination of Table 7. First, errors are reduced by the applica-
tion of the model to two years of short-term data. Second, the model applied to 1 yr of
short-term data tends to perform at least as well as estimates based on 2 yr of data
without modelling (with the exception of kinetic power density estimates at Summer-
side and Gore Bay). Third, the model estimates using 4 yr of short-term data are not
signicantly better than those made with 2 yr of monitoring, and are sometimes
marginally worse.
Finally, by considering the ve regions together, the RMS errors conrm that, in
general, the model results derived from 1 yr of short-term monitoring outperform the
estimates based on 2 yr of monitoring alone and that the model results are further
improved by using 2 yr of monitoring. At this point the RMS errors are down to less
than 4% for average wind speed estimates and less than 10% for kinetic energy
density. Additional years of monitoring do not yield further improvement in the
model estimates.
Table 8a summarizes the results of estimating long-term average wind direction at
the ve target stations based on model calculations that used 1, 2 and 4 yr of
short-term data. Comparing the last three columns in Table 8a with the long-
term observations, it was found that, except for Region E, the largest error was 15
and the next largest was 9 (see Table 8b). There was no consistent pattern of
improvement with increasing length of the monitoring period; however, the
errors in the mean directions were small enough for this not to be a concern.
Using 2 yr of observations without model application yielded estimates of the
long-term mean direction that were only in error by 14 in Regions A to C, where-
as a 33 error was obtained in Region D. Application of the model with only 1 yr
of monitoring did signicantly better (3 error) than 2 yr of monitoring alone in
Region D.
The 2 yr wind rose at Abbotsford (Region E) is highly bi-modal (see Fig. 9d), i.e.,
there is almost an equal probability that the average direction could be NNENE
or SSWSW. The fact that these two opposite directions almost balance in fre-
quency, accounts for the large errors of 140 and 147 for the model estimates based
262 J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268
Table 7
a. Long-term estimates of average wind speed SM (m s) and kinetic energy density N (W m`) at target
station compared with long- and short-term observed values. Model estimates are made for three
short-term monitoring periods
Region Observations Model estimates with monitoring period
Long term 2 yr 1 yr 2 yr 4 yr
Years SM N SM N SM N SM N SM N
A 22 5.5 173 5.2 156 5.2 151 5.4 162 5.4 164
B 21 4.4 101 4.1 82 4.5 113 4.2 93 4.2 92
C 18 4.6 104 4.9 114 4.3 83 4.7 102 4.8 108
D 21 5.3 153 5.5 169 5.1 140 5.1 142 5.5 162
E 18 2.6 32 2.8 42 2.6 38 2.7 37 2.8 38
b. Percentage errors in long-term estimates of average wind speed SM and kinetic energy density N at
target station. Observed long-term values SM (m s) and N (W m`), respectively, are copied from Table 7a.
Monitoring estimate is made from 2 yr of observations. Model estimates are made for three short-term
monitoring periods. Also shown is the RMS error
Region Years SM N SM N SM N SM N SM N
A 22 5.5 173 !5.5 !9.8 !5.5 !12.7 !1.8 !6.4 !1.8 !5.2
B 21 4.4 101 !6.8 !18.8 2.3 11.9 !4.5 !7.9 !4.5 !8.9
C 18 4.6 104 6.5 9.6 !6.5 !20.2 2.2 !1.9 4.3 3.8
D 21 5.3 153 3.8 10.5 !3.8 !8.5 !3.8 !7.2 3.8 5.9
E 18 2.6 32 7.7 31.3 0.0 18.8 3.8 15.6 7.7 18.8
RMSE 6.2 18.0 4.3 15.1 3.4 9.0 4.8 10.1
on 1- and 2 yr monitoring, respectively. For this region, therefore, it appears that
model estimates based on more than 2 yr of monitoring is needed to reduce the
direction error to an acceptable level. (With 4 yr of monitoring, for example, the error
is only 12.)
The diculty with predicting the wind direction in Region E is reected in the AE
average error and RMSE rows in Table 8b. Those values are large (approximately 30
and 60, respectively) for model estimates based on 1 or 2 yr of monitoring. For
Regions AD, on the other hand, model estimates based on only 1 yr of monitoring
tend to be more accurate than estimates based on 2 yr of monitoring without use of
the model.
The above calculations were all repeated with the incorporation of the thermal
stability class together with wind speed and direction; however, this method produced
results that were only marginally dierent from the above results using speed and
direction alone.
J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268 263
Table 8
a. Long-term estimates of circular average wind direction
M
() at target station
compared with long- and short-term observed values. Model estimates are made for
three short-term monitoring periods. Errors are with respect to the long-term ob-
served values
Region Observations Model estimates with
monitoring period
Long term 2 yr 1 yr 2 yr 4 yr
Years
M

M

M

M

M
A 22 240 241 232 234 234
B 21 289 286 292 287 286
C 18 265 269 268 250 264
D 21 231 264 234 231 222
E 18 218 189 358 071 230
b. Errors in long-term estimates of circular average wind direction
M
() at target
station
Region Years
M

M

M

M

M
A 22 240 #1 !8 !6 !6
B 21 289 !3 #3 !2 !3
C 18 265 #4 #3 !15 !1
D 21 231 #33 #3 0 !9
E 18 218 !29 #140 !147 #12
AE AVE 14 31 34 6
AD AVE 10 4 6 5
AE RMSE 20 63 66 7
AD RMSE 17 5 8 6
AVE"Average error.
RMSE"Root mean square error.
10. Summary and conclusions
The two-site correlation model of Ref. [1] has been modied to allow greater
exibility in the choice of wind speed bins and wind direction sectors.
The revised model was then tested on long-term data from ve pairs of
Canadian weather stations. The stations were chosen to cover a variety of terrain
situations.
A preliminary analysis was performed on data from each of the 10 stations to
determine the length of a short-term period of data required to estimate the long-term
264 J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268
wind climatology at the same location. It was concluded that an absolute minimum of
12 months of monitoring is needed; even so, there is still signicant uncertainty in the
estimates of the long-termaverage speed. Increasing the short-termmonitoring period
to 24 months reduces the standard deviations slightly and the worst errors more
signicantly. After 24 months, the improvement is more gradual. A short-term
monitoring period between 12 and 24 months might produce worse estimates than for
the 12-month period.
For the two-site tests, one station from each of the ve pairs was designated as
a reference station; the other was designated as a target station. Three methods of
testing were used to estimate the long-term wind speed distribution. It was found,
however, that there was very little dierence whether the correlations were based on
(i) wind speed alone; (ii) wind speed and direction; or (iii) wind speed, wind direction
and thermal stability. Similarly, method (iii) showed little dierence from method (ii)
for estimating the long-termwind direction distributions. Including the wind direction
in the correlation, however, was essential in estimating the long-term wind direction
distributions.
In general, it was found that, rst, the model results derived from 1 yr of short-
term monitoring outperform the estimates based on 2 yr of monitoring alone.
Secondly, the model results are further improved by using 2 yr of monitoring, but
additional years of monitoring do not yield further improvement in the model
estimates. The only exception to this nding was for wind direction estimates in
Region E where the target-station wind rose was highly bimodal. In that case model
estimates using more than 2 yr of monitoring were required to reduce the direction
error to acceptable levels.
Therefore, to respond to the questions raised in the Introduction: (i) a short-term
dataset at a candidate wind-turbine site must be at least 1 yr and preferably 2 yr in
length to be representative of the long-term climatology; and (ii) model results based
on 1 yr of monitoring data usually produce better long-term estimates of the wind
resource at a candidate wind-turbine site than does 1 yr of monitoring alone. The
exception to (ii) may be for wind direction estimates when the wind rose is highly
bi-modal.
For wind-energy applications, the wind speed distribution estimates produced by
the model are the most crucial results. Wind direction distribution estimates are
usually less important, but may be relevant for wind-turbine siting (e.g., avoiding
undesirable upwind roughness or turbulence generators). The wind direction esti-
mates are also important for air pollution studies.
Acknowledgements
James Salmon was supported by a contract from Atmospheric Environment Ser-
vice. John Walmsley was partially supported by the Government of Canadas Panel
on Energy Research and Development. We thank Dr. Wanmin Gong for her review
and comments on the manuscript.
J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268 265
Appendix A
Let A and B be two sites having simultaneous short-term wind data, where A is
a reference site (e.g., a weather station) with a long-term period of wind data and B is
a site at which climatological wind estimates are required. Dening d
GH
as the relative
frequency of simultaneous occurrence of wind categories i at A and j at B, then let
F
G
,
(

H
d
GH
(A.1a)
G
H
,
'

G
d
GH
(A.1b)
be the relative frequencies of occurrence of category i at A and category j at B,
respectively. The wind categories may be, for example, those of a joint distribution of
wind speed classes and wind direction bins. (For 6 classes and 16 bins, I"J"96.)
Calculate a coecient matrix c
GH
from:
c
GH
"
d
GH
F
G
(A.2)
with no summation on i. It can then be shown that
G
H
"c
GH
F
G
, (A.3)
where the summation convention is in eect. The coecient matrix c
GH
derived from
short-term simultaneous data is then used to estimate g
H
, the long-term climatology at
B from:
g
H
"c
GH
f
G
, (A.4)
where the summation convention is in eect. Here f
G
represents the long-term data at
A. Verication of the accuracy of the method can be obtained by applying Eq. (A.4) to
an independent period for which the g
H
data are available.
Walmsley and Bagg [1] presented details of the proper way to distribute cases of
calm winds amongst the wind direction bins when calculating the matrix d
GH
from the
raw data.
Appendix B
Formulae for means and standard deviations of directional data were given by
Mardia [15] and are summarized here. If

,
`
,
2
,
L
are a set of directions and P
G
is
a point on the circumference of the unit circle corresponding to
G
, then the centre of
gravity of the n points is (CM, SM), where
CM"
1
n
L

G
f
G
cos
G
, SM"
1
n
L

G
f
G
sin
G
. (B.1)
266 J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268
Here f
G
is the frequency associated with direction
G
in the case of grouped data; f
G
"1
for individual data points. The mean direction is
M"tan (SM/CM), (B.2)
which is conveniently calculated by Fortran code as M"AAN2(SM,CM) in the range
(!, ) radians.
The mean resultant length of the vectors from the origin of the unit circle to the
points P
G
is
RM"(CM`#SM`)` (B.3)
and the circular variance is
S
"
"1!RM, (B.4)
values of which lie in the range (0, 1). Finally, the circular standard deviation is
"[!2 ln(1!S
"
)]`, (B.5)
values of which lie in the range (0, R) radians.
References
[1] J.L. Walmsley, D.L. Bagg, A method of correlating wind data between two stations with application
to the Alberta Oil Sands, Atmos.-Ocean. 16 (1978) 333347.
[2] J.L. Walmsley, D.L. Bagg, Calculations of annual averaged sulphur dioxide concentrations at ground
level in the AOSERP study area, Alberta Oil Sands Environmental Research Program, Edmonton,
1977.
[3] A.D. Busse, J.R. Zimmerman, Users Guide for the Climatological Dispersion Model, Rep., EPA-R4-
73-024, US Environmental Protection Agency, Research Triangle Park, NC, 1973, pp. 131.
[4] J.L. Walmsley, Assessing the Wind Resource, in: Wind-Diesel Systems, Cambridge University Press,
Cambridge, 1994, pp. 5494.
[5] S. Kunz, M. Baumgartner, Evaluation von Extrapolationsmodellen der Windgeschwindigkeit,
OFEN, Bern, Switzerland, 1986.
[6] AWEA, Standard Procedures for Meteorological Measurements at a Potential Wind Turbine Site,
Publ. AWEA 8.1-1986. Amer. Wind Energy Assoc., Alexandria, VA, 1986, pp. 18.
[7] W.S. Kau, H.N. Lee, S.K. Kao, Statistical model for wind prediction at a mountain and valley station
near Anderson Creek, California, J. Appl. Meteorol. 21 (1982) 1821.
[8] S.R. Hanna, J.C. Chang, Representativeness of wind measurements on a mesoscale grid with station
separations of 312 m to 10 km, Boundary-Layer Meteorol. 60 (1992) 309324.
[9] I. Palomino, F. Mart n, A simple method for spatial interpolation of the wind in complex terrain,
J. Appl. Meteorol. 34 (1995) 16781693.
[10] W.R. Goodwin, G.J. McRae, J.H. Seinfeld, A comparison of interpolation methods for sparse data:
application to wind and concentration elds, J. Appl. Meteorol. 18 (1979) 761771.
[11] J.R. Salmon, J.L. Walmsley, Testing a correlation model for wind speed and direction estimates,
Proc. 2nd European and African Conf. on Wind Engineering, 2226 June 1997, Genoa, Italy,
pp. 189196.
[12] J.L. Walmsley, D. Woolridge, J.R. Salmon, MS-Micro/3 Users Guide, Rep. ARD-90-008, Atmo-
spheric Environment Service, Downsview, Ontario, Canada, 1990, 88 pp.
J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268 267
[13] I. Troen, N.G. Mortensen, E.L. Petersen, WASP Wind Atlas Analysis and Application Programme
Users Guide, Ris+ National Laboratory, Roskilde, Denmark, 1988.
[14] F. Pasquill, The estimation of the dispersion of windborne material, Meteorol. Mag. 90 (1961) 3349.
[15] K.V. Mardia, Statistics of Directional Data, Academic Press, London, 1972.
[16] J. Wieringa, P.J. Rijkoort, Windklimaat van Nederland, Staatsuitgeverij, Den Haag, 1983.
[17] N.J. Cherry, Wind energy resource survey methodology, J. Indust. Aerodyn. 5 (1980) 247280.
[18] J. Wieringa, Roughness-dependent geographical interpolation of surface wind speed averages, Quart.
J. Roy. Meteorol. Soc. 112 (1986) 867889.
268 J.R. Salmon, J.L. Walmsley/J. Wind Eng. Ind. Aerodyn. 79 (1999) 233268

Das könnte Ihnen auch gefallen