Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Journal of Wind Engineering & Industrial Aerodynamics 171 (2017) 121–136

Contents lists available at ScienceDirect

Journal of Wind Engineering & Industrial Aerodynamics


journal homepage: www.elsevier.com/locate/jweia

Estimation of roughness length at Hong Kong International Airport via


different micrometeorological methods
Y.C. He a, b, P.W. Chan c, Q.S. Li a, b, *
a
Dept. of Architecture and Civil Engineering, City University of Hong Kong, Hong Kong
b
Architecture and Civil Engineering Research Centre, City University of Hong Kong, Shen Research Institute, Shenzhen, PR China
c
Hong Kong Observatory, Kowloon, Hong Kong

A R T I C L E I N F O A B S T R A C T

Keywords: Aerodynamic roughness length scale ðz0 Þ is an essential parameter for the parameterization of momentum flux
Aerodynamic roughness length exchanges at land-atmosphere interface. In this paper, several micrometeorological methods are applied for
Wind measurement estimation of z0 based on wind measurements at Hong Kong International Airport (HKIA). The concepts of source
Micrometeorological method area and internal boundary layer are adopted to better understand the measurement results. The validity and
Wind shear alerting prediction accuracy of the estimation methods for z0 are examined and discussed. A map of terrain roughness at
HKIA is established.

1. Introduction Several methods have been developed for the estimation of z0 . These
methods can be categorized into three groups: micrometeorological (or
Surface roughness is an aerodynamic property of the earth, which is anemometric) methods (Verkaik, 2000; Powell et al., 2003; Masters
related to surface coverage, surrounding obstructions, topographic relief, et al., 2010a), classification methods (Davenport, 1960; Wieringa, 1992,
and so on. It serves continuously as a momentum sink for the atmospheric 1993), and morphometric (or geometric) methods (Lettau, 1969; Grim-
flow (Wieringa, 1993), and plays an important role in governing wind mond and Oke, 1999). Since morphometric methods are usually only
structures within the atmospheric boundary layer (ABL) (Garratt, 1994). applicable for built-up terrains, which will not be considered in
Conventionally, surface roughness can be best indexed by the aero- this study.
dynamic roughness length z0 which is regarded as an empirical measure Micrometeorological methods are driven by wind measurements.
of retarding/disturbing effects that the surface has on near-ground winds. Commonly adopted micrometeorological methods include profile
This parameter is essential for the parameterization of momentum flux method, variance method and gustiness method. Among these methods,
exchange at land-atmosphere interface, and accurate determination of its profile method requires mean speed records collected at multiple height
value has been identified as a key issue in a wide range of applications in levels, while variance and gustiness methods need turbulence measure-
wind engineering, such as determination of design wind loads on struc- ments recorded at a single level. Details about these methods will be
tures (Irwin, 2006), estimation of diffusion of pollutant plumes (Wong discussed in the following section.
and Liu, 2013), numerical simulation of environmental problems Classification methods rely on existing knowledge of z0 associated
(Blocken et al., 2007), mathematical modelling of wind field (Meng et al., with a group of basic terrain classes. The roughness length for a given
1995), assessment of wind energy potential (Emeis, 2014), etc. Surface terrain can be subjectively assessed using roughness classes and visual
roughness length is also of great importance to convert wind speeds estimation. The problem is that roughness length suggested in different
associated with different terrains, measurement heights or averaging literature for the same terrain type may vary distinctly. Wieringa (1992,
periods, and to better understanding of site-specific measurements of 1993) reviewed 30 years' roughness data from boundary-layer mea-
surface wind (Powell and Houston, 1996; Verkaik, 2000; Vickery and surements and compared 5 popular classifications of roughness. It was
Skerlj, 2005; Harper et al., 2010; Masters et al., 2010a, 2010b; Balder- found that the local-scale classification of Davenport, (1960) is reliable,
rama et al., 2011; Miller et al., 2015; He et al., 2014a, 2016). provided that the lowest two roughness classes are adjusted. Table 1 lists

* Corresponding author. Dept. of Architecture and Civil Engineering, City University of Hong Kong, Hong Kong.
E-mail address: bcqsli@cityu.edu.hk (Q.S. Li).

https://doi.org/10.1016/j.jweia.2017.09.019
Received 8 May 2017; Received in revised form 21 September 2017; Accepted 24 September 2017

0167-6105/© 2017 Elsevier Ltd. All rights reserved.


Y.C. He et al. Journal of Wind Engineering & Industrial Aerodynamics 171 (2017) 121–136

Table 1 introduces the estimation methods considered in this paper. Section 3


Davenport classification of effective terrain roughness. describes the observation sites and datasets of wind measurements. The
Class Landscape description z0 ðmÞ estimation results are presented in Section 4. Conclusions and main
1 Sea Open sea or lake (irrespective of wave size), tidal flat, 0.0002 findings of this study are summarized in Section 5.
snow-covered flat plain, featureless desert, tarmac
and concrete, with a free fetch of several kilometers 2. Introduction of estimation methods
2 Smooth Featureless land surfaces without any noticeable 0.005
obstacles and with negligible vegetation; e.g.,
beaches, pack ice without large ridges, marsh, and
2.1. Source area and internal boundary layer
snow-covered or fallow open country.
3 Open Level country with low vegetation (e.g., grass) and 0.03 Due to adoption of wind measurements, z0 derived via micrometeo-
isolated obstacles with separations of at least 50 rological methods depends on both surface covers and arrangement of
obstacle heights; e.g., grazing land without
measurement systems. It is stressed that a device placed above a site
windbreaks, heather, moor and tundra, runway area
of airport. Ice with ridges across-wind. explores only a portion of its surroundings. Meteorologically, the portion
4 Roughly Cultivated or natural area with low crops or plant 0.10 of upstream surface that contains the effective sources contributing to the
open covers, or moderately open country with occasional flux exchanges at the concerned site is termed as the source area (or
obstacles (e.g., low hedges, isolated low buildings or
footprint). A large source area accounts for a large-scale average of sur-
trees) at relative horizontal distances of at least 20
obstacle heights
face properties. The source area is a function of observation height, sta-
5 Rough Cultivated or natural area with high crops or crops of 0.25 bility condition, and surface roughness (Schmid and Oke, 1990). For
varying heights, and scattered obstacles at relative wind measurement, it is elliptical in shape and is aligned in the upwind
distances of 12–15 obstacle heights for porous objects direction (~30 in width) from the concerned site (WMO, 2008). Under
(e.g., shelterbelts) or 8 to 12 obstacle heights for low
neutral condition the area with significant contribution to measurements
solid objects (e.g., buildings) (analysis may need zd )
6 Very rough Intensively cultivated landscape with many rather 0.5 atop a typical mast lies upwind at a distance of several kilometers, while
large obstacle groups (large farms, clumps of forest) increasing instability reduces the source area to a region closer to the
separated by open spaces of about 8 obstacle heights. concerned site.
Low densely-planted major vegetation like bushland, In principle, terrain classifications are established for uniform ter-
orchards, young forest. Also, area moderately
covered by low buildings with interspaces of 3–7
rains. In reality, however, it is not uncommon to encounter terrain
building heights and no high trees (analysis requires changes in an area at a scale larger than several hundreds of meters.
zd ) Under such heterogeneous conditions, wind structures are dominated by
7 Skimming Landscape regularly covered with similar-size large 1.0 both local (or new) surface and upwind (initial) exposure. Internal
obstacles, with open spaces of the same order of
boundary layer (IBL) will be formed at the interacting area where at-
magnitude as obstacle height; e.g., mature regular
forests, densely built-up area without much building mospheres gradually adapt to the new surface (Garratt, 1990). The IBL
height variation (analysis requires zd ) depth hI varies with the distance downwind of fetch change x (Powell
8 Chaotic City centers with mixture of low-rise and high-rise 2 and Houston, 1996):
buildings, or large forests of irregular height with
many clearings (analysis by wind tunnel advised)
hI ¼ cz0R ðx=z0R Þ0:8 (1)

where c (0.28–0.75) is a constant that depends on stability status


the updated version of Davenport's classification (Davenport et al., 2000) (c  0:3 under neutral condition), z0R is the larger value of roughness
that is recommended by World Meteorological Organization (WMO, length of new and initial fetches. Wind structures above hI are governed
2008). One may refer to Stewart and Oke (2012) for further information by upwind surface cover, while those below 0:1hI are completely
to better understand the Davenport's classification. adjusted to the new terrain. Wind flows in the middle region demon-
Given the important role of z0 in governing the ABL and the diversity strate a blended feature that is influenced by both initial and new ter-
of estimation methods for z0 , there is a need to assess the prediction rains: while small eddies are dominated by the immediate action of
performance of different approaches. However, such works especially local obstructions, large eddies exist in accordance with large-scale
those concerning varied micrometeorological methods are limited, as terrain setups.
wind measurements required for such comparisons are rarely available. For an area with evident variation of terrain setups, the effective
Among few related studies, Barthelmie et al. (1993) compared the pre- roughness length, i.e., the roughness length producing a representative
diction results by using classification method, profile method, gustiness momentum flux for the concerned area (Fielder and Panofsky, 1972),
method and speed-variance method. Considerable method-related vari- may be estimated by analyzing wind measurements recorded at the
ations in z0 were reported, with estimations in the same azimuth sector blending height, i.e., the height at which the flow is approximately in
varying by a factor of as much as 20. Verkaik and Holtslag (2007) equilibrium with the local surface and also independent of horizontal
analyzed the roughness length using three micrometeorological methods. position (Mason, 1988). Alternatively, it may be estimated by averaging
It was reported that under inhomogeneous conditions, the profile method lnðz0i Þ (called the lnðz0 Þ avearage method hereafter), z0i being the
might be invalid and estimation results of z0 via different roughness length of each small-scale homogeneous patch (WMO, 2008).
wind-turbulence based methods differed evidently. It is further note- But the effective roughness length tends to be larger than the average of
worthy that most estimation methods for z0 may suffer from significant z0 values of all patches, due to the fact that it is easier to charge the at-
uncertainty. Verkaik (2000) evaluated the performance of two gustiness mosphere with turbulence than to discharge it by dissipation (Wieringa,
methods and found that different parametric settings in the models could 1993). Thus, the values of z0 listed in Table 1 constitute a lower limit of
result in great discrepancy of estimation results. effective roughness that can occur in terrain situations where such sur-
The motivation of this study is twofold: First, to assess the perfor- face cover is dominant.
mance of several roughness estimation methods through cross compari-
son analysis of the results from these methods; Second, to explore the 2.2. Wind profile based method
characteristics of surface wind at Hong Kong International Airport
(HKIA) and determine associated roughness length so as to advance wind Within the surface layer of neutrally stratified atmosphere, vertical
shear alerting for aircraft operation at the airport. profiles of horizontal mean wind speed can be depicted by the logarith-
The remainder of this paper is organized as follows. Section 2 mic law:

122
Y.C. He et al. Journal of Wind Engineering & Industrial Aerodynamics 171 (2017) 121–136

u* z Similarly, for wind direction, one has:


UðzÞ ¼ ln (2)
κ z0 κ cv
σd ¼ (6)
in which, κ  0:4 is the von Karman constant, u* is the surface friction lnðz=z0 Þ
velocity, z denotes the height above the zero-displacement plane, or
where σ u and σ d (unit: rad) are the standard deviation of wind speed and
z ¼ z0  d, z0 and d being the height above ground and the zero-
wind direction, respectively; cu ; cv are coefficients depending upon sta-
displacement, respectively. Under open flat terrain condition, d can be
bility condition. Under neutral condition, cu ¼ 2:2, cv ¼ 1:9 for unfiltered
regarded as zero. But for rough surfaces (i.e., forests and built-up terrain)
measurements of σ u and σ d , while for filtered measurements in which σ u
where wake-interference or skimming flows exist, d should be taken into
is expected to be attenuated by about 12%, and σ d by 2%, cu ¼ 1:94, cv ¼
account so as to obtain reliable estimation of z0 .
1:86 (WMO, 2008). Fig. 1 presents the dependence of z0 on σ u and σ d at
Thus, given mean wind speeds U1 ; U2 at two heights of z1 ; z2 , z0 can
varied height z and with different inputting values of cu and cv .
be computed by:

U1 lnðz2 Þ  U2 lnðz1 Þ 2.4. Gustiness method


lnðz0 Þ ¼ (3)
U1  U2
As a technique driven by measurements of gust factor (GF), the
When wind records at multi-height levels are available, z0 can be
gustiness method has several forms that are characterized by adopted GF
determined as the intercept of the wind profile curve in the altitude co-
model (e.g., Wieringa, 1973; Beljaars, 1987). This study mainly considers
ordinate through fitting.
the method documented by Masters et al. (2010a,b). According to the
definition of GF, one has:
2.3. Variance method
Uτ;max Uτ;max  UT σ u σu
GF≡ ¼1þ ⋅ ¼ 1 þ gðT; τ; zÞ⋅ (7)
The best way of estimating z0 recommended by WMO (2008) is the UT σu UT UT
variance method. According to the similarity theory, the normalized
standard deviation of horizontal wind speed by the surface friction ve- where τ and T denote respectively gust duration and mean speed dura-
locity is a function of stability: tion, and gðT; τ; zÞ is the peak factor which may be expressed as
(Davenport, 1964):
σ u =u* ¼ cu ðz=L; h=LÞ (4) pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi .pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
gðT; τ; zÞ ¼ 2 lnðυTÞ þ 0:5772 2 lnðυTÞ (8)
where L is the Monin-Obukhov length, and h is the ABL depth.
Substituting Equation (2) into the above equation results in:
in which υ is the zero up-crossing rate. Equation (8) is established based
κ cu on the assumption of Gaussian distribution of wind speed records. When
σ u =U ¼ (5)
lnðz=z0 Þ non-Gaussianity of wind records should be taken into account (Balder-
rama et al., 2012), other revised forms, such as that proposed by Kareem

Fig. 1. Dependence of roughness length on variance of wind turbulence.

123
Y.C. He et al. Journal of Wind Engineering & Industrial Aerodynamics 171 (2017) 121–136

σ 2u ðT; τ; z; z0 Þ ¼ ∫ 0 Su ðn; z; z0 Þχ 2 ðnÞdn (15)
and Zhao (1994), may be adopted.
In reality, measurements of σ u and gðT; τ; zÞ are inevitably attenuated In Masters et al. (2010a), the von Karman wind spectrum was adopted
by measurement systems. For a typical anemometry system chain con- to determine the filtered wind speed variance σ u ðT; τ; z; z0 Þ:
sisting of anemometer, transmission line and recorder, spectrum of 
recorded wind turbulence Sr ðnÞ may be expressed as (Beljaars, 1987; n⋅Su ðnÞ 4nLxu U
¼h   2 i5=6 (16)
Verkaik, 2000): σu
1 þ 70:8 nLxu U
Sr ðnÞ ¼ SðnÞ⋅χ 2 ðnÞ (9)
in which turbulence integral length Lxu can be calculated by a combined
χ 2 ðnÞ ¼ T1 ðnÞ⋅Thp ðnÞ⋅Tra ðnÞ (10) usage of Taylor's hypothesis, i.e., Lxu ¼ UTu , and an empirical estimator of
the integral time scale recommended in Engineering Sciences Data Unit
 2 1
(ESDU, 1983), i.e., Tu ¼ 3:13z0:2 (ESDU, 1983), while σ u is the unfiltered
T1 ðnÞ ¼ 1 þ ð2πnλ=UÞ (11) wind variance formulated by the Harris variance model (ESDU, 1985):

 1 16

Thp ðnÞ ¼ 1  1 þ ð2πnTÞ2 (12) 7:5u* η½0:538 þ 0:09 lnðz=z0 Þη


σ u ðzÞ ¼ ; η ¼ 1  6 f z=u* (17)
1 þ 0:156 ln½u* =jfz0 j
 2  2
sinðπnt0 Þ sinðπnΔNÞ The effective roughness length can be computed using an iterative
Tra ðnÞ ¼ ; or Tra ðnÞ ¼ (13)
πnt0 πnΔ method based on the above equations. Fig. 2 shows correlations between
GF and z0 estimated via the above introduced gustiness method, by
Here, SðnÞ is the unfiltered power spectrum, Tl account for low-pass considering the effects of four key inputting parameters, i.e., τ, T, z and λ.
filtering effect due to the inertia of anemometer's rotating components, Without specific notification, the defaulted parameter values involved in
Thp represents high-pass filter associated with the measuring period T, Tra each figure are set as: τ ¼ 3s, T ¼ 600s, U600 ¼ 10m=s, z ¼ 10m, λ ¼ 5m.
stands for running average filtering effects for the analog (first) or
discrete (second) signals, λ is the distance constant of the anemometer, n 2.5. Marine method
is the frequency (unit: Hz), t0 is the analog running-average duration, Δ
and N are the sampling interval and sample number of instantaneous Owing to the mobility of seawater, the state of sea surface, and
readings involved in the discrete average. therefore the marine roughness, depends greatly on wind strength. Other
υ and σ u can be expressed as: factors, such as water depth, can also influence the marine roughness
evidently (Smith et al., 1992). The Charnock model is conventionally
∞ adopted to estimate the roughness length over waters (Charnock, 1955;
∫ 0 n2 Su ðn; z; z0 Þχ 2 ðnÞdn
υ2 ¼ ∞ (14) Garratt, 1977):
∫ 0 Su ðn; z; z0 Þχ 2 ðnÞdn

Fig. 2. Dependence of z0 on varied inputting parameters via gustiness method.

124
Y.C. He et al. Journal of Wind Engineering & Industrial Aerodynamics 171 (2017) 121–136

  pffiffiffiffiffiffiffiffiffiffi
z0 ¼ αu2* g (18) z0 ¼ 10 exp  κ Cd;10 (22)

where α is a constant in range of 0.015–0.035 (Powell et al., 2003). Given in which Cd;10 ¼ u2* =U10
2
is the drag coefficient at 10 m AMSL, which may
the value of α, z0 can be determined using iterative method based on be empirically determined by (Smith et al., 1992):
Equations 2 and 18.
Fairall et al. (2003) examined the results from a number of mea- 103 Cd;10 ¼ 0:27 þ 0:116U10 (23)
surement programs and recommended the following speed-dependent
The above equation was derived based on field measurements at a
form of α:
coastal area where the local water depth was 18 m. Other formulas of
8
< 0:011 for U10  10 Cd;10 can be found in Vickery and Skerlj (2000), Harper et al. (2010) and
α ¼ 0:011 þ 0:000875ðU10  10Þ for 10 < U10 < 18 (19) He et al. (2016).
:
0:018 for 18  U10  25 The above models predict that the marine roughness length increases
monotonously with wind speed. However, some wind profile observa-
Oost et al. (2002) proposed the following an exponential relationship tions over deep oceans (Powell et al., 2003; Vickery et al., 2009) revealed
between α and wave age: that the drag coefficient, and therefore the roughness length, may level
  2:5  off or even reduce when wind speed exceeds a certain value. Thus, it is
α ¼ 50 Cp u* ; Cp ¼ gTp ð2πÞ (20) reasonable to use the capped forms of these estimators. For simplifica-
tion, it may be regarded that Cd;10 derives the maximum value when
where Cp and Tp are respectively the phase speed and period of the U10 ¼ 25 m=s based on the results reported by Vickery et al. (2009).
dominant wave, and Cp =u* is a measure of the wave age. For fully Fig. 3 shows dependence of z0 on mean wind speed using the capped
developed waves in deep waters, Tp ¼ 0:729U10n , U10n being the neutral forms of the above introduced four marine methods.
wind speed at 10 m above mean sea level (AMSL).
Unlike the above studies which focused on modelling of α, Taylor and 3. Introduction of observation sites and datasets
Yelland (2001) modelled the marine roughness directly as a function of
the significant wave height and peak wavelength: 3.1. HKIA
  .
4:5
z0 ¼ 1200Hs Hs Lp ; Lp ¼ gTp2 ð2πÞ (21) As an international metropolis, Hong Kong has one of the busiest
airports in the world. As is well known, low-level wind shear could be
where Hs is the significant wave height and Lp is the wavelength hazardous to departing/arriving aircrafts at the airport. Previous studies
associated with the dominant wave period (thus, Hs =Lp is approximately have shown that airflow disturbances induced by topography/terrain and
the slope of the dominant wave). For fully developed artificial structures (such as airport terminal) are a major cause of sig-
waves, Hs ¼ 0:0248U10n
2
. nificant low-level wind shear at HKIA (Liu et al., 2010; Chan and Lee,
Another kind of estimation method for marine roughness focuses on 2011; Chan, 2012a, 2012b; Li and Chan, 2012). To improve the safety of
modeling the drag coefficient which has a relationship with z0 as follows: airport operations, great efforts have been made to further understanding
of low-level wind shear around HKIA, with a focus on utilizing numerical

Fig. 3. Dependence of marine roughness on mean wind speed.

125
Y.C. He et al. Journal of Wind Engineering & Industrial Aerodynamics 171 (2017) 121–136

techniques to facilitate timely forecasting of such wind phenomena. meters to 12 m), besides parking aprons with frequently berthed air-
Although insightful results have been achieved, some basic assumptions planes. To protect the airport against sea waves and currents, a groyne
involved in the numerical studies are needed to be validated. One of them was constructed on the north side of HKIA. The slope of the groyne ex-
is the value setting of roughness length which not only governs ceeds 0.3.
approaching wind fields during the simulations but also plays an essen-
tial role in treatment of boundary layer condition. 3.2. Equipment and datasets
The geographical location of HKIA is shown in Fig. 4. It is located at
Pearl River Estuary, China. The northern and western parts of the airport HKIA has been equipped with various meteorological instruments at
are surrounded by seawater, while complicated terrain in Lantau Island different sites. As depicted in Fig. 4, seven masts were erected alone two
(with peaks rising to ~1000 m AMSL, and valleys ~400 m AMSL in runways: A1E, R1E, R1C, R1W, R2E, R2C and R2W (“R1” and “R2” stand
between) exist to the south of HKIA. Inside the airport, there are two for the south runway and the north runway, while “E” “C” and “W”
runways: the north and the south. Both runways have an orientation of denote east, central and west, respectively). Each mast has been equipped
70/250 . The area in the vicinity of the runways is covered by low grass, with anemometers and vanes at 10 m height above ground level (AGL) to
that to the south of the runway corridors is occupied by a number of measure speed and direction of horizontal wind component. The ane-
freight depots whose heights are in a range of several meters to over mometers are of cup type (Thies 4.3303 model), whose distance constant
20 m, while between them there are airport terminal (at eastern part, is 5 m. The detection accuracies of the anemometer and vane are 0.3 m/s
maximum building height is ~30 m), midfield concourse building (at and 1 , respectively.
middle part, maximum building height is ~30 m), and hangars and Two kinds of datasets from the masts are analyzed in this study:
warehouses (at western part, building heights are in a range of several routine measurements and fine records. The routine measurements

Fig. 4. Location of the study sites and surrounding terrain features.

126
Y.C. He et al. Journal of Wind Engineering & Industrial Aerodynamics 171 (2017) 121–136

include 1 min mean wind speed and direction as well as 3 s peak gust in records are further tested according to the following two criteria to
each minute, which were recorded every minute based on instantaneous exclude those collected under either rainy or non-neutral conditions: (1)
readings that are sampled at 1 Hz. Ten years' (from January 2005 to The horizontal component of mean surface-layer speed  6 m/s. (2) The
December 2014) data of this kind is available for this study. The fine amplitude of vertical component of mean surface-layer speed  3 m/s.
records consist of 1 s–1 s updated wind speed and direction that were Throughout this study, the surface layer denotes the portion of the ABL
collected during a period of 44 days between July 2006 and July 2012 below 100 m AGL.
when strong wind warning signals (for both tropical cyclones and mon-
soons) were issued by Hong Kong Observatory (HKO). Except R1E, the 4. Results and discussions
fine records are available for this study.
Besides the above introduced mechanical devices, A1E has also been 4.1. Comparison of results around A1E
equipped with a Doppler sonic-detection-and-ranging (Sodar) profiler
(located at several meters far away from the meteorological mast). The 4.1.1. Wind profile and z0;p
profiler (Aerovironment Model 4000) operates at a central acoustic fre- Measurements at A1E are analyzed first to investigate the wind
quency of 4.5 kHz. It probes three components of wind by emitting three characteristics around HKIA, and particularly to compare the estimation
acoustic beams (i.e., a vertical beam and two oblique beams at about 15 performance for z0 by different micrometeorological methods. Owing to
from the vertical) at 0.33 Hz. Mean wind velocity vectors at 39 gate levels unavailability of the refine records from the mast at A1E, those collected
from 10 m to 200 m AGL were recorded every minute based on ensemble at R1E (distance between R1E and A1E is 270 m) are utilized
backscatter signals associated with each gate. A 3 years' (2010–2012) substitutively.
remote sensing dataset from the profiler is available for this study. Fig. 5 shows the ensemble-mean profiles of horizontal wind speed
All the above three kinds of datasets are utilized in this study. The normalized by the mean surface-layer speed. The results are stratified
records from the profiler are analyzed to obtain the vertical profiles of into 12 azimuth sectors, according to the mean surface-layer direction.
mean horizontal wind speed so that the roughness length can be deter- The number of qualified profiles (N) in each sector is depicted in the
mined via the profile method. The measurements from the masts are figure. As can be seen, influenced by nearby buildings and surrounding
analyzed to investigate turbulent characteristics of wind near ground. topography, mean wind profiles in most sectors do not follow the loga-
The fine records can be used to estimate z0 via both the gustiness method rithmic law in the surface layer. In the northern and southern sectors,
and the variance method. On the other hand, the long-term routine there are evident profile kinks around 30–40 m. But the wind profiles in
measurements allow us to examine the dependence of estimation results two western sectors (i.e., 240 –270 and 270 –300 ) associated with an
on sample size and historical time through comparison analysis. open-flat exposure are distributed logarithmically.
Based on the above observations, wind profiles in Sector 250 –290
3.3. Quality control of datasets are selected to represent the profiles above open terrain, as shown in
Fig. 6. Two fitting lines are drawn in the figure via the least square
To guarantee the quality of the wind measurements adopted in this method based on the gate records in the range of 15–50 m (fit-1) and
study, two kinds of tests are conducted on the raw datasets from the 50–100 m (fit-2), respectively. The estimated values of z0 , denoted as z0;p ,
masts, which aim to examine the thermal stability and stationarity of the are 0.05 m (fit-1) and 0.003 m (fit-2), with 95% confidence bounds being
wind records, respectively (He et al., 2013). Measurements which fail to 0.007–0.1 m and 0.0006–0.006 m, respectively. The results basically
pass the tests are discarded in the following analysis. Due to absence of agree with the recommendations for Class-3 (z0 ¼ 0.03 m) and Class-2 (z0
weather information required for the qualification of stratification status, ¼ 0.005 m) in the terrain classification listed in Table 1. The dependence
only neutrally stratified winds are concerned in this study. Practically, it of z0;p on different fitting ranges indicates the existence of IBL in this
is regarded that neutral condition can be satisfied in strong wind cases. sector (i.e., land area in Sector 250 –290 ) where is associated with a
Measurement results documented in Wieringa (1973) suggested that smooth-to-rough terrain change.
neutral condition can be achieved when mean wind speed exceeds a From Fig. 4, the land area in Sector 250 –290 is featured by two
threshold level of about 5.5 m/s for daytime and 5.0 m/s for nighttime. In patches of terrain cover: open terrain dominated by runway and low
this study, it is assumed that neutral condition can be achieved when grass in an area of ~1 km upwind the study site, and an area featured by
10 min mean wind speed exceeds 6 m/s. occasional large obstacles beyond the above. If the values of z0 for these
Tests for stationarity of the wind speed records are conducted patches are assumed to be 0.03 m and 0.1 m in accordance with the
following the way adopted in He et al. (2013) which includes four re- suggestion in Table 1, the effective roughness length at this area
quirements for each 10 min data run: (1) Standard deviation of wind computed via the ln z0 avearage method turns to be 0.055 m, which is
direction  15 . (2) Difference between scalar and vector mean almost identical to z0;p derived on the basis of the gate records below
speed  0.51 m/s (or 1 knot). (3) Peak wind speed should not be beyond 50 m height. It is then assumed that approaching wind flows experience a
5 times standard deviations away from mean. (4) Time series of wind step change (i.e., marine-to-land) in underlying terrains before reaching
speed should pass tests for 1st order stationarity at a significant level of to the observation site, with z0 before and after the terrain transition
0.05 using the reverse arrangement method. being 0.0002 m (Class-1 in Table 1) and 0.055 m, respectively. Using
This study does not examine the validity of remote sensing mea- Equation (1) (distance downwind of fetch change  3 km), the IBL depth
surements from the profiler, since relevant studies were conducted by is calculated as 100 m. According to the IBL theory, wind flows between
Chan (2008). The quality control for this dataset aims to examine the 50 and 100 m above the observation site are influenced by both the
reliability of data at each gate and the completeness of credible gate overland terrain and marine exposure. Thus, it is reasonable that the
records for each wind profile. It is regarded that a gate record is credible measured z0;p for this range is associated with a Class-2 terrain in Table 1.
if associated signal-to-noise ratio (SNR)  8 (He et al., 2014b), and a
wind profile is complete if the number of credible gate records contained 4.1.2. GF and z0;GF
in the range of 15–100 m exceeds 14 (there are totally 18 gates in this Values of GF vary with adopted gust duration τ and mean speed
range). Here, the lowest gate records (i.e., at 10 m height) are discarded duration T. In practice, the ratio between 3 s peak gust and 10 min mean
because interference effect at this level (due to surrounding obstacles) is wind speed, i.e., GFð3;600Þ , is most frequently used (WMO, 2008).
generally severe, whilst gate records at a height above 100 m are dis- Following this convention, 3 s peak gust over 10 min duration is
regarded since the logarithmic law is only applicable in the surface layer computed using overlapping average technique to get GFð3;600Þ for each
whose depth is in the order of 100 m in moderate-to-strong wind cases. 10 min wind record.
The selected profile measurements consisting only of credible gate

127
Y.C. He et al. Journal of Wind Engineering & Industrial Aerodynamics 171 (2017) 121–136

Fig. 5. Vertical profiles of normalized longitudinal mean wind speed at A1E.

Fig. 7 presents 25th, 50th, and 75th percentiles of GFð3;600Þ and asso- 0 –60 and Sector 150 –360 . The maximum relative difference of me-
ciated values of z0 estimated via the gustiness method ðz0;GF Þ, based on dian of GF in these sectors is 2.5%. In Sector 240 –300 which corre-
the routine measurements at A1E (A1E-R) and R1E (R1E-R) and the fine sponds to a flat exposure, GF values from different datasets almost
records at R1E (R1E-F). Statistics of the obtained results are not discussed coincide with each other, reflecting the credibility of the GF results. The
in terms of mean and standard deviation, as they often demonstrate non- median of z0;GF in Sector 240 –270 and Sector 270 –300 is ~0.002 m
Gaussian feature (He et al., 2013; Lombardo and Krupar, 2017). The and ~0.02 m, different from ~0.05 m obtained via the profile method
results are stratified into 12 azimuth sectors according to the mean wind based on the lower portion of profiles in Sector 250 –290 . From Fig. 4,
direction ðθÞ. Data records involved in each sector are generally large the source area in Sector 240 –270 is dominated by the south runway
enough in size to make the statistics results meaningful. corridor, while that in Sector 270 –300 is additionally influenced by the
The results at A1E-R and R1E-R agree considerably well in Sector midfield concourse building. Therefore, it is reasonable that these two
sectors respectively belong to Class-2 (featured by “no obstacles”) and
Class-3 (“few isolated obstacles”) stipulated in the terrain classification.
The noticeable difference between z0 estimations via the gustiness
method and the profile method can be explained by the height-
dependence of source area above the observation site. As discussed
previously, the hangars and warehouses located at the western part of
HKIA can influence the flux exchanging process above A1E in a vertical
range of 10–100 m (i.e., 0.1–1hI ). However, below this range, incident
atmospheres are completely adjusted to the local terrain.
In Sector 60 –150 , the results at A1E-R and R1E-R show distinct
discrepancy. Winds measured at A1E are consistently “gustier” than
those detected at R1E. Since A1E is located closer to upwind hills and
civil structures, it is believed that atmospheres over A1E are more
influenced by the disturbing effects of upwind obstacles than those over
R1E. Actually, previous studies of field measurements and numerical
simulations showed that flow separation and vortex may exist downwind
of obstacles (Chan, 2012b; He et al., 2014b). Such wind-turbulence
complex tends to be dissipated during the transferring process both
downwind of the obstacle and downward from the generation point,
resulting in obstacle-relative dependence of GF (He et al., 2013). How-
ever, it is unclear why this dependence of GF becomes less evident in
Fig. 6. Wind profiles and z0;p at A1E. Sector 150 –210 where the topographic effects should be even more

128
Y.C. He et al. Journal of Wind Engineering & Industrial Aerodynamics 171 (2017) 121–136

Fig. 7. Measurements of GF and z0;GF at A1E and R1E.

significant. A possible explanation is that the observation sites are located difference between σ u =U and σ d in sectors except 150 –180 and
just downwind of the main part of Lantau Island (Fig. 4) in this sector, 270 –330 are less than 10%. Comparatively large differences in Sector
and the gradient of GF around A1E and R1E decreases to a level at which 150 –180 and Sector 270 –330 are speculated to correlate with the
the distance between the two masts becomes insufficiently long discrepancy of responding features of anemometers and vanes in non-
to resolve. equilibrium atmospheres downwind of the mountain or the
Comparison of the results at R1E-R and R1E-F reveals an interesting airport terminal.
phenomenon: under hilly exposure conditions GFs from the fine records To compare the estimation performance of the gustiness method and
are consistently larger than those from the routine measurements, while the variance method, results of z0;GF based on the fine records at R1E are
the above discrepancy vanishes in other sectors associated with less also depicted in Fig. 9. It is observed that while z0;σu and z0;GF agree well
evident topography. To explore this phenomenon, Fig. 8 examines the under relatively rough terrain conditions (roughly z0 > 0:03 m), z0;σu
relationship between GF, z0;GF and mean wind speed, based on the becomes consistently larger than z0;GF under smoother terrain conditions.
routine measurements at A1E and R1E in two selected sectors (i.e., As pointed out by Verkaik and Holtslag (2007), the largest eddies of at-
60 –120 and 120 –210 ). As reflected, GF and z0;GF generally increase mosphere are those that adapt slowest of all to terrain changes. Because
with mean wind speed. Because the refine records were collected during σ u in the surface layer is mostly determined by the largest eddies, z0;σu is
strong wind events, GF and z0;GF in these sectors at R1E-F are reasonably expected to better represent the large-scale average of surface roughness.
larger than those at R1E-R. The speed-dependence of wind turbulence By contrast, peak gust speed, and therefore GF, depends more signifi-
downwind of hills may be attributed to two reasons: First, flow separa- cantly on local surface properties. Thus, the discrepancies between z0;σu
tion, leeward wave/vortex and other topography-induced phenomena, and z0;GF reflect the rough-to-smooth transition in terrain around the
are more likely to occur under strong winds; second, when such phe- observation site in associated sectors (Fig. 4).
nomena occur, the behavior of wind turbulence, such as the frequency of
vortex shedding, is closely related to Reynolds number of approach at- 4.2. Results with marine exposure
mosphere or approach wind speed. It is emphasized that the mean wind
speed records depicted in Fig. 8 were collected downwind of Lantau Is- The speed dependence of GFð3;600Þ and z0 estimated via the gustiness
land where shielding effect is evident. Thus, the speed-dependence of GF
method and the four marine methods that are respectively depicted by
and z0;GF may only be revealed partially.
Equations (19)–(21), (23) is examined in Fig. 10, based on the routine
measurements in Sector 270 –330 at R2E, R2C and R2W. Values of
4.1.3. Variance of wind turbulence and z0;σ
inputting variables for Equations (19)–(21), (23) are set in accordance
The fine records at R1E are analyzed to derive the standard deviation
with the information of the measurement systems at the study sites. The
of wind speed σ u and wind direction σ d via the technique documented by
results are stratified into 5 speed groups with 3 m/s intervals. The
Yamartino (1984). Results of 25th, 50th, and 75th percentiles of GFð3;600Þ
number of data runs involved in each group is marked in the figure. For
and associated z0 values estimated via the variance method ðz0;σu ; z0;σd Þ z0 , measurements at the three sites are composited to enlarge the sample
are presented in Fig. 9. size in strong-wind groups. Since the observation sites are located only
The good agreement between the results of σ u and σ d demonstrates several meters away from coastal line, wind measurements may be
the validity of the collected fine records. The maximum relative influenced non-negligibly by the topographic effects caused by the

Fig. 8. Speed dependence of GF and z0;GF at A1E and R1E.

129
Y.C. He et al. Journal of Wind Engineering & Industrial Aerodynamics 171 (2017) 121–136

Fig. 9. Measurements of wind variance ðσ u ; σ d Þ and z0;σ at R1E.

Fig. 10. Speed-dependence of GF and marine z0 based on routine measurements.

groyne. Using the method stated in European Standard for wind actions predictions from the marine roughness models. The above findings sug-
on structures (EN, 1991-1-4, 2005; Ngo and Letchford, 2008, 2009), the gest that the adopted gustiness method is able to provide reasonable
orography factor for mean wind speed at the observation sites is esti- estimations of z0 under marine exposure conditions, although the
mated as co;U  1:09 which means a decrease by 8.3% in measured GF if method is conventionally not utilized in marine environment.
topographic effects on peak gust are neglected. If the orography factor for The results based on the fine records at the three sites in Sector
peak gust is assumed to be co;bu  1:04, there will be a 4.6% decrease in 270 –330 are listed in Table 2. Limited by sample size, no classification
measured GFs. The results of z0;GF based on the raw dataset (medians of speed group is considered herein. Values of σ u =U, σ d and GF estimated
depicted by blue squares) and the records after corrections by consid- both with (“w”) and without (“wo”) correction of the topographic effects
ering only co;U (black circles) and both co;U and co;bu , or co;GF ¼ co;bu =co;U  on wind measurements are documented. Following the previous discus-
sions, GF is corrected by using co;GF , while σ u is corrected via
1:05 (pink triangles), are presented in the figure.
co;σu  1 þ ðco;bu  1Þ=g  1:01, with g  3 being the peak factor. Thus,
As demonstrated, the estimation results are extremely sensitive to the
topographic effects. While values of z0;GF based on the raw dataset at this the orography factor for σ u =U turns to be co;TI  co;σu =co;U  0:93. The
coastal area are unreasonably (almost one order of magnitude) smaller orography factor for σ d is assumed as the same as co;TI .
than those predicted by the marine roughness models that were estab- Similar to the results based on the routine measurements, the medians
lished based on observations over open oceans, those from corrected GF of z0;GF without correction for the topographic effects are about one order
measurements by using co;U are several times larger than the model of magnitude smaller than those predicted by Equation (23), while the
predictions. But the results after correction for the topographic effects on values based on the corrected measurements agree well with the model
both mean speed and peak gust show good agreement with the predictions. Values of z0;σu and z0;σd become almost doubled after the

Table 2
Estimations of marine z0 via varied methods based on fine records.

Site Sector 270 –330 25th, 50th, 75th percentiles of turbulence 25th, 50th, 75th percentiles of z0 z0 estimated via marine methods
measurements estimations

No. of runs U (m/s) θ ( ) σu/U σd (wo/w) GF z0,σu z0,σd z0,GF z0,Eq19 z0,Eq20 z0,Eq21 z0,Eq23
(wo/w) (rad) (wo/w) (mm) (mm) (mm) (mm) (mm) (mm) (mm)

R2E 64 10.8 306 0.078/.084 0.070/.065 1.18/1.24 0.48/1.0 0.24/.51 0.001/.062 0.19 0.13 0.48 0.35
0.091/.098 0.080/.074 1.21/1.27 2.0/3.7 0.91/1.8 0.011/.28
0.096/.104 0.097/.090 1.26/1.32 3.1/5.6 4.7/8.0 0.18/1.9
R2C 60 11.0 306 0.082/.089 0.069/.064 1.19/1.25 0.78/1.6 0.21/.44 0.003/.011 0.20 0.14 0.50 0.38
0.091/.098 0.080/.074 1.23/1.29 2.0/3.7 0.91/1.8 0.037/.66
0.104/.112 0.094/.087 1.25/1.31 5.7/10 3.7/6.4 0.11/1.4
R2W 72 10.8 309 0.085/.092 0.070/.065 1.20/1.26 1.1/2.1 0.24/.51 0.005/.18 0.19 0.13 0.48 0.35
0.094/.102 0.082/.076 1.23/1.29 2.6/4.8 1.1/2.2 0.037/.66
0.106/.114 0.103/.095 1.28/1.34 6.6/11 7.3/12 0.42/3.6

130
Y.C. He et al.
Table 3
Measurement results along airport corridors.

Site Sector-1 Sector-2


(50 –80 at R1E, R1C, R1W; 70 –100 at R2E, R2C, R2W) (240 –270 at A1E, R1C, R1W; 220 –250 at R2E, R2C, R2W)

No. U (m/s) 25th, 50th, 75th percentiles of turbulence 25th, 50th, 75th percentiles of z0 No. of U (m/s) 25th, 50th, 75th percentiles 25th, 50th, 75th percentiles of z0 estimation
of runs & θ( ) measurements estimation runs & θ( ) of turbulence measurements

σu/U σd GF z0,σu z0,σd z0,GF σu/U σd GF z0,σu z0,σd z0,GF


(rad) (mm) (mm) (mm) (rad) (mm) (mm) (mm)

R1E F 245 9.7 0.119 0.110 1.31 14 11 1.3 93 10 0.100 0.091 1.25 4.4 2.8 0.10
67 0.132 0.128 1.36 28 30 5.3 250 0.114 0.106 1.30 11 8.6 1.1
0.150 0.154 1.43 57 79 23 0.130 0.130 1.36 25 33 5.6
R 5685 7.6 – – 1.25 – – 0.13 2430 7.2 – – 1.27 – – 0.33
73 1.30 1.2 250 1.32 1.9
1.38 9.2 1.38 10
A1E R 5352 7.2 – – 1.32 – – 1.8 1361 7.0 – – 1.27 – – 0.33
71 1.37 7.6 252 1.32 1.8
1.45 33 1.38 10
R1C F 386 9.4 0.121 0.110 1.31 16 11 1.4 49 11 0.106 0.095 1.26 6.5 4.0 0.17
66 0.135 0.131 1.37 32 35 6.7 254 0.117 0.121 1.29 13 21 0.75
0.150 0.156 1.44 57 86 28 0.135 0.146 1.40 31 61 12
R 14,956 7.7 – – 1.26 – – 0.24 1619 7.0 – – 1.25 – – 0.18
71 1.31 1.6 250 1.30 1.3
131

1.38 8.6 1.37 7.7


R1W F 400 9.7 0.107 0.096 1.28 7.2 4.1 0.37 54 11 0.091 0.074 1.21 2.0 0.40 0.05
66 0.120 0.111 1.33 16 12 2.6 251 0.105 0.087 1.25 6.3 1.9 0.12
0.136 0.129 1.40 34 31 15 0.132 0.111 1.30 28 12 0.84
R 12,749 7.7 – – 1.25 – – 0.16 1551 7.0 – – 1.20 – – 0.011
70 1.30 1.2 248 1.24 0.12
1.37 7.9 1.30 1.5

Journal of Wind Engineering & Industrial Aerodynamics 171 (2017) 121–136


R2E F 318 12 0.124 0.094 1.30 19 3.5 0.83 405 11 0.108 0.096 1.27 7.7 4.1 0.29
87 0.136 0.104 1.37 33 7.8 3.8 229 0.120 0.108 1.32 15 10 1.8
0.150 0.119 1.40 57 19 15 0.137 0.123 1.38 35 24 8.2
R 20,385 7.7 – – 1.25 – – 0.18 6504 8.0 – – 1.29 – – 0.70
93 1.30 1.0 231 1.35 4.3
1.35 4.8 1.42 21
R2C F 386 12 0.124 0.090 1.30 19 2.6 1.1 441 11 0.103 0.087 1.26 5.5 1.9 0.14
86 0.137 0.101 1.35 34 6.3 4.2 229 0.114 0.097 1.29 11 4.7 0.62
0.151 0.116 1.41 58 17 17 0.129 0.110 1.35 25 11 3.8
R 24,106 7.9 – – 1.27 – – 0.31 9795 7.9 – – 1.24 – – 0.11
92 1.31 1.5 232 1.30 0.87
1.37 6.7 1.36 5.8
R2W F 383 12 0.126 0.092 1.31 21 3.0 1.1 321 11 0.092 0.069 1.22 2.1 0.22 0.014
86 0.139 0.102 1.36 37 6.9 4.9 227 0.105 0.080 1.26 6.1 0.93 0.13
0.151 0.114 1.42 58 15 18 0.119 0.092 1.31 14 3.0 1.2
R 32,008 8.1 – – 1.28 – – 0.50 7927 7.7 – – 1.20 – – 0.012
90 1.33 2.4 228 1.25 0.12
1.38 10 1.30 0.96
Y.C. He et al. Journal of Wind Engineering & Industrial Aerodynamics 171 (2017) 121–136

correction process, but both kinds of results are several times larger than From Fig. 4, there is an artificial island located to the northeast of
those predicted by Equation (23). As the best estimation of local marine HKIA. This island was constructed during 2012–2014. Based on the
roughness is unavailable, the validity and prediction accuracy of the above discussions, wind flows at the observation heights above R1E and
variance method need to be further examined. A1E in Sector-1 experienced a marine-to-land change in terrain before
the appearance of the island, while they should undergo multi-step
4.3. Results along airport corridors changes in terrain afterward. To examine the possible effects of the
artificial island on the measurements, Fig. 11 plots the evolutions of
Both routine measurements (“R”) and fine records (“F”) in two sectors GFð3;600Þ in Sector-1 at both A1E and R1E (“Aver-1” represents GFð3;600Þ
at all the seven study sites are analyzed to investigate turbulent wind for each data run, “Aver-50” stands for mean values of every 50 runs,
characteristics and roughness length along the two airport corridors. The while “Aver-all” is the ensemble-mean value). As demonstrated, no
results are summarized in Table 3. For R1E, R1C and R1W (i.e., sites evident variation trend can be detected from time history of GFð3;600Þ .
along the south runway), the two azimuth sectors are 50 –80 and Actually, the above issue featured by multi-step changes in upwind ter-
240 –270 , while for R2E, R2C and R2W (i.e., sites along the north rains may be simplified according to the recommendations in ESDU
runway), the two azimuth sectors are 70 –100 and 220 –250 . (1982) which suggests that the intermediate island-occupied area inside
These results reveal the following features: the marine region can be ignored since the span of the island is less than
that of the marine area between the island and HKIA.
(1) Values of z0;GF are consistently smaller than those estimated via
the variance method. The reasons are twofold. First, wind- 4.4. Results downwind of obstacle zones
variance measurements carry more information of large-scale
eddies. Thus, wind indexed by σ appears to be gustier than that To examine turbulent wind characteristics and surface roughness in
indexed by GF. Second, as to be discussed later, the estimation areas downwind of some typical obstacle zones in HKIA, field mea-
results tend to be more sensitive to model coefficients under surements in two azimuth sectors are selected: Sector-1, i.e., Sector
smooth terrain condition, and significantly increased un- 330 –360 at R1E, A1E and R1W and 0 -30 at R1C, and Sector-2, i.e.,
certainties may also account for the distinct discrepancy. Sector 150 –180 at R2E and R2W and Sector 120 –150 at R2C. From
(2) Estimations of z0;GF based on the fine records are noticeably larger Fig. 4, Sector-1 at R1W and Sector-2 at R2W are associated with an
than those derived via the routine measurements in Sector-1 (with upwind region occupied by scattered buildings and parking aprons.
hilly exposure), but the difference tends to decrease with the Sector-1 at R1E and A1E and Sector-2 at R2E are associated with the
distance downwind of the topography, which respectively sug- terminal building, whilst Sector-1 at R1C and Sector-2 at R2C corre-
gests the dependence of GF and z0;GF on wind strength and hill- spond to the midfield concourse building. Here, measurements in two
relative position. However, values of z0;GF estimated through the sectors around each of the three obstacle zones are analyzed to offer
two kinds of datasets in Sector-2 (with open flat exposure) differ insights for the validity of the obtained results. Measurements at R1E
insignificantly. Meanwhile, estimations of z0 via different and A1E in Sector-1 may further manifest the dependence of wind
methods in Sector-2 are consistently smaller than their counter- turbulence on the distance downwind of the terminal building through
parts in Sector-1. The above observations reveal the topographic comparison analysis.
effects on the obtained results. The obtained results are summarized in Table 4. Basically, the results
(3) Results of z0;GF in Section-2 reveal a marine-to-land terrain change in Sector-1 agree well with their counterparts in Sector-2, taking into
along each of the two runways. The medians of z0;GF increase account the topographic effects and other uncertain factors. In most
gradually from ~0.1 mm at the western part, to ~0.8 mm at the cases, median of z0 is in the order of 0.2 m. Again, z0;GF derived from the
middle part, and then to ~2 mm at eastern part, reflecting the fine records is found to be typically larger than that based on the routine
progressively strengthened influence of local terrain on the mea- measurements. As mentioned previously, turbulent wind structure
surement results. If the local terrain is assumed to belong to Class- downwind of a large obstacle depends greatly on associated Reynolds
2 (z0 ¼ 5 mm) of the terrain classification in Table 1, the IBL number, or incident wind strength and characteristic dimension of the
depths at these three parts turn to be 70 m, 140 m, and 220 m, obstacle. On the one hand, GF results, obtained particularly downwind of
respectively. It is clear that the observation points at R1W and the terminal building, reflect that stronger winds are prone to result in
R2W are located in the blending range of the IBL where atmo- more turbulent flows at the observation sites. On the other hand, com-
spheres are influenced by both the marine exposure and the local parison of the results at R1E with those at A1E reveals the distance
terrain, whilst those at R1E and R2E are located inside the lowest dependence of GF downwind of the terminal building. The above find-
one-tenth portion of the IBL where wind structures are in equi- ings suggest that the disturbing influence of big airport buildings on
librium with the local surface. Similar results are also observed departing/arriving aircrafts should be investigated in terms of wind
along the north runway in Section-1. strength and building-relative position.

Fig. 11. Evolutions of GF in Sector 50 –80 at two sites.

132
Y.C. He et al.
Table 4
Measurement results downwind of obstacle zones.

Site Sector-1 Sector-2


(330 –360 at R1E, A1E, R1W; 0 –30 at R1C) (150 –180 at R2E, R2W; 120 –150 at R2C)

No. of runs U (m/s) & θ( ) 25th, 50th, 75th percentiles 25th, 50th, 75th percentiles No. of runs U (m/s) & θ( ) 25th, 50th, 75th 25th, 50th, 75th percentiles of z0 estimation
of turbulence measurements of z0 estimation percentiles of turbulence
measurements

σu/U σd GF z0,σu z0,σd z0,GF σu/U σd GF z0,σu z0,σd z0,GF


(rad) (m) (m) (m) (rad) (m) (m) (m)

R1E or R2E F 125 9.3 0.160 0.161 1.45 0.079 0.099 0.036 200 8.5 0.170 0.175 1.49 0.10 0.14 0.059
345 0.181 0.185 1.53 0.14 0.18 0.093 166 0.187 0.199 1.60 0.16 0.24 0.18
0.198 0.211 1.62 0.20 0.30 0.22 0.216 0.231 1.70 0.27 0.40 0.38
R 859 7.2 – – 1.37 – – 0.008 1614 7.2 – – 1.41 – – 0.016
343 1.44 0.028 167 1.52 0.089
133

1.52 0.087 1.66 0.29


A1E R 797 7.4 – – 1.35 – – 0.004 – – – – – – – –
344 1.40 0.015
1.49 0.058
R1C or R2C F 238 7.4 0.154 0.152 1.46 0.065 0.074 0.038 356 10.0 0.146 0.148 1.39 0.049 0.065 0.011
16 0.172 0.181 1.57 0.11 0.16 0.14 131 0.162 0.166 1.46 0.083 0.11 0.037

Journal of Wind Engineering & Industrial Aerodynamics 171 (2017) 121–136


0.194 0.215 1.72 0.18 0.31 0.43 0.184 0.195 1.54 0.15 0.22 0.11
R 1011 6.9 – – 1.39 – – 0.012 4616 7.6 – – 1.31 – – 0.0014
17 1.47 0.048 131 1.37 0.0071
1.59 0.17 1.45 0.032
R1W or R2W F 91 9.8 0.176 0.199 1.49 0.12 0.24 0.061 165 9.6 0.166 0.169 1.47 0.092 0.12 0.043
346 0.203 0.213 1.61 0.22 0.31 0.20 163 0.184 0.193 1.55 0.15 0.21 0.11
0.221 0.230 1.70 0.30 0.39 0.37 0.207 0.223 1.65 0.24 0.35 0.26
R 977 7.1 – – 1.34 – – 0.004 1055 8.1 – – 1.47 – – 0.047
341 1.41 0.018 166 1.56 0.14
1.52 0.082 1.68 0.33
Y.C. He et al. Journal of Wind Engineering & Industrial Aerodynamics 171 (2017) 121–136

measurements to extract records with evident nonneutral/nonstationary


features, the thermal effects and unanticipated noise interference may
still exist in some “qualified” specimens. From Fig. 8, the GF values in
Sector 60 –120 at A1E and R1E demonstrate a decreasing trend with the
increase of wind speed for moderate-to-strong winds, which is usually
regarded as a feature relevant to the thermal effect (Shu et al., 2015).
Meanwhile, field measurements stratified in a certain azimuth sector may
contain records influenced by different kinds of exposure, although great
attention has been made during the stratification process. The way of
azimuthal stratification may result in evident uncertainty in estimations
of z0;p . It is shown that estimations of z0;p in Sector 240 –270 and Sector
270 –300 based on the lower portion of wind profiles (below 50 m) are
0.2 m and 0.02 m, respectively. However, if stricter criteria for data
quality control and/or finer group intervals for data stratification are
adopted, there would be insufficient specimens left in the concerned
sectors, due to the characteristics of regional climate and local topo-
Fig. 12. Roughness map at HKIA. graphic features.
Second, the results from both the fine records and the routine
4.5. Uncertainty analysis measurments demonstrate that the mean GF values are typically 1%–3%
larger than associated medians. Accordingly, z0;GF estimated via mean of
The uncertainty sources can be classified into two groups which are GF is larger than that estimated via median of GF by a factor of 0.2–1. For
related to z0 estimation model and dataset, respectively. Model-related σ u and σ d , the mean value generally differs from the median by a factor of
uncertainty consists of potential errors due to inaccurate value assign- 1%–5%, which is associated with a difference between z0;p and z0;σu by a
ment of some key model parameters, while data-related uncertainty is factor of 0.2–0.8. Such uncertainties related to adopted statistical pa-
largely owing to the way of data selection/sratification and data analysis rameters have been discussed by Lombardo and Krupar (2017).
(such as fitting process). Third, the estimation of z0;p is sensitive to the portion of adopted wind
From Figs. 1–2, the model predictions via either the varance or profiles for data fitting. Different observation heights correspond to
gustiness methods are considerably sensitive to a number of parameters, varied source areas. Thus, narrow range of wind profiles is preferred
especially under smooth terrain condition. From Fig. 1, given σ u =U ¼ during fitting process from the footprint point of view. However, nar-
0:10 at z ¼ 10 m which is consistent with the measurement results with a rowing the fitting range will decrease the number of gate records, and
marine exposure as listed in Table 2, z0;σu computed using different therefore, the accuracy of fitting results. It is suggested by Wieringa
values of cu can vary from 0.4 mm ðcu ¼ 2:5Þ to 6 mm ðcu ¼ 1:9Þ and to (1993) that the least profile levels needed for determination of z0;p within
300 mm ðcu ¼ 1:5Þ. Actually, for simplification purpose, it is sometimes a factor of 2 are 3 for high roughness (z0 ¼ 1 m), 4 for moderate
assumed that cu ¼ 2:5 (thus κ⋅cu  1 in Equation (5)) in practice even for roughness (z0 ¼ 0:1 m), and 5 or more over really smooth terrain.
filtered measurements (cu should be equal to 1.94). The above simplifi-
cation may result in an estimation error up to an order of magnitude in
z0;σu . For the gustiness estimation method, the parameters of gust period, 4.6. Roughness map at HKIA
average wind duration, observation height and distance constant can
affect the estimation results significantly (Fig. 2). Among these parame- A recommended roughness map at HKIA is shown in Fig. 12. The
ters, the distance constant λ, which serves as an index for the dynamic marine roughness length is ~0.5 mm based on the results listed in Table 2
features of an anemometer (Kristensen and Hansen, 2002), may not be which is associated with a fixed incident wind speed of 11 m/s. This
determined accurately. It is also possible that λ varied with historical value is slightly smaller than 0.2 mm for “open-sea” terrain suggested in
service/maintain conditions of the anemometer system. From Fig. 2, the terrain classification (Table 1). But it is identical to that recom-
given GFð3;600Þ (based on the field measurements in Table 2), z0;GF mended by Harper et al. (2010) for the “at-sea” exposure class for tropical
calculated by using different λ values varies from 0.17 mm ðλ ¼ 0Þ to cyclone forecasting purposes. For the runway corridors, the preferred
0.40 mm ðλ ¼ 10Þ and to 1.0 mm ðλ ¼ 20Þ. For smaller GF values, the estimation of z0 is 5 mm, based on a comprehensive consideration of the
difference between estimations of z0;GF via varied λ values should be even measured results listed in Table 3. This value corresponds to the
more significant. Other parameters involved in the estimation model for “smooth” terrain class in Table 1.
z0;GF , such as spectrum of wind turbulence and turbulence integral length The area to the south of HKIA and the western part of the region
scale, may also bring in non-negligible uncertainties for estimation re- between the two runway corridors are dominated by scatter obstacles,
sults especially under rugged-exposure conditions. such as buildings and berthed airplanes, and z0 for both areas is sug-
For data-related uncertainties, three aspects are stressed herein. First, gested to be 0.3 m, based on the results in Table 4. This value is associ-
although a series of data quality tests have been conducted on raw ated with the “rough” landscape class as suggested in the terrain
classification. The effective roughness length for the open area to the
Table 5 west of the midfield concourse building is estimated as 0.03 m, based on
Comparison of z0 results at HKIA from different data sources. the measured results at R1C in Sector 300 –330 . This value corresponds
Surface Type z0 (m)
to an “open” class in the terrain classification. Results around the ter-
minal building and the midfield concourse building are believed to
Liu et al. Leung et al. Li and Chan This study
depend greatly on building-relative location and wind strength. Thus, no
(2010) (2012) (2012, 2016)
specific values of z0 are presented in Fig. 12.
Sea 0.0002 0.015 0.0005 0.0005
By considering the above results, the large-scale average of z0 at A1E
Asphalt 0.0024 0.0015 – –
Runway 0.004 0.00015 – 0.005
in the western sector is calculated as 0.04 m via the lnðz0 Þ average
Building wall 0.02 0.0015 – – method. This value approximates to the measured z0;p (0.05 m) based on
Short grass 0.01 0.015 – – the lower portion of wind profile as shown in Fig. 6. This result is also
Airport island – – 0.005 0.05 consistent to the conventional assumption that an airport area belongs to
Lantau Island – – 0.5 0.5
an “open” terrain with z0  0:03 m.

134
Y.C. He et al. Journal of Wind Engineering & Industrial Aerodynamics 171 (2017) 121–136

Table 5 compares the values of z0 obtained in this study with smaller than z0;σu or z0;σd in case with a rough-to-smooth terrain
those assumed in four computational studies on wind characteristics at transition.
near ground level around HKIA. Estimations of z0 for “Lantau Island” in (4) The results of vertical wind profiles fail to provide the information
this study is derived based on the measurement results in Sector on z0;p under rugged terrain conditions, as the profiles are dis-
120 –210 at sites along the south runway. The measured median of z0 torted severely by nearby topography. Under flat exposure con-
from the fine records in this sector is in a range of 0.2–0.5 m, compared to ditions, estimation of z0;p depends greatly on the fitting range of
that of 0.1–0.3 m based on the routine measurements. Taking into ac- wind profile, due to the discrepancy of source area. The z0;p value
count the mountain-relative effect, z0 for “Lantau Island” is adjusted to be derived on the basis of lower portion of wind profile is consistent
0.5 m which corresponds to the “very rough” terrain in Table 1. It is with the effective roughness length estimated via the wind tur-
emphasized that the above z0 value should not be used to determine the bulence measurements.
vertical profile of horizontal wind speed, since the vertical wind profiles (5) Wind measurements collected at sites located close to coastal lines
can be distorted severely downwind of hills/mountains. may be influenced by the topographic effects of groyne. The re-
In light of the fact that the runway corridors consist mainly of asphalt sults after correction for such topographic effects show that the
roads and low grass, values of z0 for “asphalt” and “short grass” may be roughness length estimated via the gustiness method is compa-
estimated via the classification method by further considering the rable to the predictions from the four marine roughness models.
measured results. According to the terrain classification in Table 1, However, values of z0;σ are noticeably larger than those predicted
“tarmac and concrete with a free fetch of several kilometers” belongs to by the models.
Class-1 (z0 ¼ 0:0002 m), while “low vegetation, e.g., grass” belongs to (6) Wind structures downwind of huge obstacles at HKIA, e.g., ter-
Class-3 (z0 ¼ 0:03 m). Thus, it is reasonable to assume that the roughness minal building, are influenced by both incident wind strength and
values for these two terrain types at HKIA are 0:01  0:02 (“short grass”) observation position. Thus, case studies via either numerical
and 0.002 m (“asphalt”). It is interesting that the effective z0 for the simulations or wind tunnel tests are needed to analyze the wind
runway corridors is calculated to be 0.0055 m using the lnðz0 Þ average characteristics in the wake region of such obstacles.
method by assuming z0 ¼ 0:015 m and z0 ¼ 0:002 m for “short grass” (7) There are a number of uncertain factors involved in the obtained
and “asphalt”. This value agrees well with the recommended value of z0 results. Of particular importance are accurate determination of
0.005 m in Table 1. key coefficients involved in each estimation model for z0 and
reasonable selection and sectionalization of datasets. The model
coefficients may not be simplified arbitrarily under smooth terrain
5. Concluding remarks conditions, since the estimation results become considerably
sensitive to these variables, such as cu , for small z0 . It is also
Aerodynamic roughness length is an essential parameter for the suggested that medians of GF or σ should be adopted for analysis
parameterization of surface wind. It plays an important role in many of terrain roughness.
aspects of engineering practice. Thus, accurately estimating z0 is of great (8) A roughness map at HKIA was established. Values of z0 obtained
practical significance. However, limited comparative studies have been in this study are compared with those assumed in previous
reported to examine the validity and estimation accuracy of various simulation works. The results of this study provide useful infor-
micrometeorological methods, possibly due to unavailability of sufficient mation for further numerical studies to advance wind shear
field measurements of winds. alerting for aircraft operation at HKIA.
Based on long-term wind measurements from both anemometers
and a Sodar wind profiler system at seven meteorological stations at In light of the important role of z0 in governing the ABL structure, it is
KHIA, this paper investigated the characteristics of surface wind and essential to provide accurate estimation of z0 to avoid excessive ampli-
directional roughness length at HKIA. Several estimation techniques for fication of z0 -related estimation uncertainties and errors due to potential
z0 , including the profile method, variance method for both wind speed propagations in subsequent analysis. Therefore, there is a need to
and wind direction, the gustiness method, classification method, and perform further studies to examine the validity and accuracy of different
four marine methods were adopted in this study. The validity of micrometeorological estimation methods for roughness length under
collected field measurements and the estimation performances of these different terrain conditions.
prediction techniques for z0 were examined through comprehensive
comparison analyses. The concept of source area and the IBL theory
Acknowledgment
were adopted to explain the obtained results. A roughness map at HKIA
was finally drawn. Main findings and conclusions of this study are
This work is fully supported by a grant from the National Natural
summarized below.
Science Foundation of China (Project No.: 51408520), a grant from the
Research Grants Council of Hong Kong Special Administrative Region,
(1) Under flat exposure conditions, the results of GFð3;600Þ and z0;GF
China (Project No: CityU 118213), a grant from Environment and Con-
derived from the long-term routine measurements agree well with
servation Fund of Hong Kong (Project No: 9211097 (19/2015)) and a
those from the fine records collected during a much shorter
grant from Fundamental Research Program of Shenzhen Municipality
period. But the two kinds of results differ evidently in sectors with
(Project No. JCYJ20160401100137854).
rugged exposures, owing to the speed-dependence of wind char-
acteristics downwind of hills. Atmospheres in stronger wind
events are consistently “gustier”. References
(2) Basically, the results of σ u =U and z0;σu agree well with their
Balderrama, J.A., Masters, F.J., Gurley, K.R., Prevatt, D.O., Aponte-Bermudez, L.D.,
counterparts of σ d and z0;σd , suggesting the validity of wind vari- Reinhold, T.A., Pinelli, J.P., Subramanian, C.S., Schiff, S.D., Chowdhury, A.G., 2011.
ance measurements and the rationality of using wind variance The Florida coastal monitoring program (FCMP): a review. J. Wind Eng. Ind.
Aerodyn. 99, 979–995.
records to estimate roughness length.
Balderrama, J.A., Masters, F.J., Gurley, K.R., 2012. Peak factor estimation in hurricane
(3) Under heterogeneous terrain conditions, surface wind flows are in surface winds. J. Wind Eng. Ind. Aerodyn. 102, 1–13.
non-equilibrium status. Winds at lower heights are more influ- Barthelmie, R.J., Palutikof, J.P., Davies, T.D., 1993. Estimation of sector roughness
enced by local surface cover than those at higher altitudes. At the lengths and the effect of prediction of the vertical wind speed profile. Boundary-Layer
Meteorol. 66, 19–47.
observation height, the results reflect that wind variance accounts Beljaars, A.C.M., 1987. The influence of sampling and filtering on measured wind gusts.
for a larger source area than that of GF, and z0;GF tends to be J. Atmos. Ocean. Technol. 4, 613–626.

135
Y.C. He et al. Journal of Wind Engineering & Industrial Aerodynamics 171 (2017) 121–136

Blocken, B., Stathopoulos, T., Carmeliet, J., 2007. CFD simulation of the atmospheric Li, L., Chan, P.W., 2016. LIDAR observation and numerical simulation of vortex/wave
boundary layer: wall function problems. Atmos. Environ. 41, 238–252. shedding at the eastern runway corridor of the Hong Kong International Airport.
Chan, P.W., 2008. Measurement of turbulence intensity profile by a mini-sodar. Meteorol. Meteorol. Appl. 23, 379–388.
Appl. 15, 249–258. Liu, C.H., Leung, D.Y.C., Man, A.C.S., Chan, P.W., 2010. Computational fluid dynamics
Chan, P.W., 2012a. A significant wind shear event leading to aircraft diversion at the simulation of the wind flow over an airport terminal building. J. Zhejiang University-
Hong Kong international airport. Meteorol. Appl. 19, 10–16. Science A 11, 389–401.
Chan, P.W., 2012b. Observation and numerical simulation of an event of vortex/wave Lombardo, F.T., Krupar, R.J., 2017. Characterization and comparison of aerodynamic
shedding from a mountain near the Hong Kong International Airport. Meteorol. Appl. roughness lengths using ground-based photography and sonic anemometry. J. Struct.
19, 371–384. Eng. 04017049.
Chan, P.W., Lee, Y.F., 2011. Application of a ground-based, multi-channel microwave Mason, P.J., 1988. The formation of areally-averaged roughness lengths. Q. J. R.
radiometer to the alerting of low-level windshear at an airport. Meteorol. Z. 20, Meteorol. Soc. 114, 399–420.
423–429. Masters, F.J., Vickery, P.J., Bacon, P., Rappaport, E.N., 2010a. Toward objective,
Charnock, H., 1955. Wind stress on a water surface. Q. J. R. Meteorol. Soc. 81, 639–640. standardized intensity estimates from surface wind speed observations. Bull. Am.
Davenport, A.G., 1960. Rationale for determining design wind velocities. J. Struct. Div. Meteorol. Soc. 91, 1665–1681.
Am. Soc. Civ. Eng. 86, 39–68. Masters, F.J., Tieleman, H.W., Balderrama, J.A., 2010b. Surface wind measurements in
Davenport, A.G., 1964. Note on the distribution of the largest value of a random function three Gulf Coast hurricanes of 2005. J. Wind Eng. Ind. Aerodyn. 98, 533–547.
with application to gust loading. Proc. Instit. Civ. Eng. 28, 187–196. Meng, Y., Matsui, M., Hibi, K., 1995. An analytical model for simulation of the wind field
Davenport, A.G., Grimmond, C.S., Oke, T.R., Wieringa, J., 2000. Estimating the roughness in a typhoon boundary layer. J. Wind Eng. Ind. Aerodyn. 56, 291–310.
of cities and sheltered country. In: Preprints of the Twelfth American Meteorological Miller, C., Balderrama, J.A., Masters, F., 2015. Aspects of observed gust factors in
Society Conference on Applied Climatology (Asheville, NC, United States), pp. 96–99. landfalling tropical cyclones: gust components, terrain, and upstream fetch effects.
Emeis, S., 2014. Current issues in wind energy meteorology. Meteorol. Appl. 21, Boundary-Layer Meteorol. 155, 129–155.
803–819. Ngo, T.T., Letchford, C.W., 2008. A comparison of topographic effects on gust wind speed.
EN 1991-1-4, 2005. Eurocode 1: Actions on Structures – Part 1-4: General Actions – Wind J. Wind Eng. Ind. Aerodyn. 96, 2273–2293.
Actions. The European Union Per Regulation 305/2011, Directive 98/34/EC, Ngo, T.T., Letchford, C.W., 2009. Experimental study of topographic effects on gust wind
Directive 2004/18/EC. speed. J. Wind Eng. Ind. Aerodyn. 97, 426–438.
ESDU, 1982. Strong winds in the Atmospheric Boundary Layer Part 1: Hourly-mean Wind Oost, W.A., Komen, G.J., Jacobs, C.M., Oort, C.V., 2002. New evidence for a relation
Speeds. Engineering Sciences Data Unit 82026, p. 61. between wind stress and wave age from measurements during ASGAMAGE.
ESDU, 1983. Strong winds in the Atmospheric Boundary Layer Part 2: Discrete Gust Boundary-Layer Meteorol. 103, 409–438.
Speeds. Engineering Sciences Data Unit 83045, p. 37. Powell, M.D., Houston, S.M., 1996. Hurricane Andrew's landfall in south Florida. Part I:
ESDU, 1985. Characteristics of Atmospheric Turbulence Near the Ground Part Π: Single standardizing measurements for documentation of surface wind fields. Weather
Point Data for Strong Winds. Engineering Sciences Data Unit 85020, p. 42. Forecast. 11, 304–328.
Fairall, C.W., Bradley, E.E., Hare, J.E., Grachev, A.A., Edson, J.B., 2003. Bulk Powell, M.D., Vickery, P.J., Reinhold, T.A., 2003. Reduced drag coefficient for high wind
parameterization of sir-sea fluxes: updates and verification for the COARE algorithm. speeds in tropical cyclones. Nature 422, 279–283.
J. Clim. 16, 571–591. Schmid, H.P., Oke, T.R., 1990. A model to estimate the source area contributing to
Fielder, F., Panofsky, G., 1972. The geostrophic drag coefficient and the effective turbulent exchange in the surface layer over patchy terrain. Q. J. R. Meteorol. Soc.
roughness length. Q. J. R. Meteorol. Soc. 98, 213–220. 116, 965–988.
Garratt, J.R., 1977. Review of drag coefficients over oceans and continents. Mon. Weather Shu, Z.R., Li, Q.S., He, Y.C., Chan, P.W., 2015. Gust factor for tropical cyclone, monsoon
Rev. 105, 915–929. and thunderstorm winds. J. Wind Eng. Ind. Aerodyn. 142, 1–14.
Garratt, J.R., 1990. The internal boundary layer-a review. Boundary-Layer Meteorol. 50, Smith, S.D., Anderson, R.J., Oost, W.A., Kraan, C., Maat, N., Decosmo, J., Katsaros, K.B.,
171–203. Davidson, K.L., Bumke, K., Hasse, L., Chadwick, H.M., 1992. Sea surface wind stress
Garratt, J.R., 1994. The Atmospheric Boundary Layer. Cambridge University Press. and drag coefficients: the HEXOS results. Boundary-Layer Meteorol. 60, 109–142.
Grimmond, C.S.B., Oke, T.R., 1999. Aerodynamic properties of urban areas derived from Stewart, I.D., Oke, T.R., 2012. Local climate zones for urban temperature studies. Bull.
analysis of surface from. J. Appl. Meteorol. 38, 1262–1292. Am. Meteorol. Soc. 93, 1879–1900.
Harper, B.A., Kepert, J.D., Ginger, J.D., 2010. Guidelines for Converting between Various Taylor, P.K., Yelland, M.J., 2001. The dependence of sea surface roughness on the height
Wind Averaging Periods in Tropical Cyclone Conditions. World Meteorological and steepness of the waves. J. Phys. Oceanogr. 31, 572–590.
Organization WMO/TD-No. 1555. Verkaik, J.W., 2000. Evaluation of two gustiness models for exposure correction
He, Y.C., Chan, P.W., Li, Q.S., 2013. Wind characteristics over different terrains. J. Wind calculations. J. Appl. Meteorol. 39, 1613–1626.
Eng. Ind. Aerodyn. 131, 51–69. Verkaik, J.W., Holtslag, A.A.M., 2007. Wind profiles, momentum fluxes and roughness
He, Y.C., Chan, P.W., Li, Q.S., 2014a. Standardization of raw wind speed data under lengths at Cabauw revisited. Boundary-Layer Meteorol. 122, 701–719.
complex terrain conditions: a data-driven scheme. J. Wind Eng. Ind. Aerodyn. 131, Vickery, P.J., Skerlj, P.F., 2000. Elimination of exposure D along hurricane coastline in
12–30. ASCE7. J. Struct. Eng. 126, 545–549.
He, Y.C., Chan, P.W., Li, Q.S., 2014b. Field measurements of wind characteristics over Vickery, P.J., Skerlj, P., 2005. Hurricane gust factors revisited. J. Struct. Eng. 131,
hilly terrain within surface layer. Wind Struct. 19, 541–563. 825–832.
He, Y.C., Chan, P.W., Li, Q.S., 2016. Standardization of offshore surface wind speeds. Vickery, P.J., Wadhera, D., Powell, M.D., Chen, Y., 2009. A hurricane boundary layer and
J. Appl. Meteorol. Climatol. AMS 55, 1107–1121. wind field model for use in engineering applications. Appl. Meteorol. Climatol. 48,
Irwin, P.A., 2006. Exposure categories and transitions for design wind loads. J. Struct. 381–405.
Eng. ASCE 132, 1755–1763. Wieringa, J., 1973. Gust factors over open water and built-up country. Boundary-layer
Kareem, A., Zhao, J., 1994. Analysis of non-Gaussian surge response of tension leg Meteorol. 3, 424–441.
platforms under wind loads. J. Offshore Mech. Arct. Eng. 116, 137–144. Wieringa, J., 1992. Updating the Davenport roughness classification. J. Wind Eng. Ind.
Kristensen, L., Hansen, O.F., 2002. Distance Constant of the RisØ Cup Anemometer. Aerodyn. 41–44, 357–368.
Technical Report RisØ-R-1320(EN). RisØ National Laboratory, Roskilde, Denmark. Wieringa, J., 1993. Representative roughness parameters for homogeneous terrain.
Lettau, H., 1969. Note on aerodynamic roughness-parameter estimation on the basis of Boundary-Layer Meteorol. 63, 323–363.
roughness-element description. J. Appl. Meteorol. 8, 828–832. Wong, C.C.C., Liu, C.H., 2013. Pollutant plume dispersion in the atmospheric boundary
Leung, D.Y.C., Lo, W.Y., Chow, W.Y., Chan, P.W., 2012. Effect of terrain and building layer over idealized urban roughness. Boundary-Layer Meteorol. 147, 281–300.
structures on the airflow in an airport. J. Zhejiang Univ. Sci. A 13, 461–468. World Meteorological Organization, 2008. Guide to Meteorological Instruments and
Li, L., Chan, P.W., 2012. Numerical simulation study of the effect of buildings and Methods of Observation. WMO-No.8, seventh ed., p. 681
complex terrain on the low-level winds at an airport in typhoon situation. Meteorol. Yamartino, R.J., 1984. A comparison of several ‘single-pass’ estimators of the standard
Z. 21, 183–192. deviation of wind direction. J. Clim. Appl. Meteorol. 23, 1362–1366.

136

You might also like