You are on page 1of 8

Renewable Energy 89 (2016) 598e605

Contents lists available at ScienceDirect

Renewable Energy
journal homepage: www.elsevier.com/locate/renene

A new methodology for urban wind resource assessment


~es, Ana Estanqueiro*
Teresa Simo
rio Nacional de Energia e Geologia e LNEG, Estrada do Pao do Lumiar, 1649-038 Lisboa, Portugal
Laborato

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 2 March 2015
Received in revised form
10 August 2015
Accepted 6 December 2015
Available online 24 December 2015

In the latest years the wind energy sector experienced an exponential growth all over the world. What
started as a deployment of onshore projects, soon moved to offshore and, more recently to the urban
environment within the context of smart cities and renewable micro-generation. However, urban wind
projects using micro turbines do not have enough prot margins to enable the setup of comprehensive
and expensive measurement campaigns, a standard procedure for the deployment of large wind parks.
To respond to the wind assessment needs of the future smart cities a new and simple methodology for
urban wind resource assessment was developed. This methodology is based on the construction of a
surface involving a built area in order to estimate the wind potential by treating it as very complex
orography. This is a straightforward methodology that allows estimating the sustainable urban wind
potential, being suitable to map the urban wind resource in large areas. The methodology was applied to
a case study and the results enabled the wind potential assessment of a large urban area being consistent
with experimental data obtained in the case study area, with maximum deviations of the order of 10%
(mean wind speed) and 20% (power density).
2015 Elsevier Ltd. All rights reserved.

Keywords:
Urban wind
Resource assessment
Micro-generation
CFD

1. Introduction
Urban wind energy has a large potential to be explored in the
context of smart cities, whether through the installation of small
wind turbines in the domestic sector (building rooftops and surrounding areas), or integrated in the building envelope providing
that they are designed with wind energy exploitation in mind [1].
The wind potential in urban areas is difcult to characterize due to
the high impact of obstacles and structures on the atmospheric
ow. Buildings often cause ow separation, wind speed reduction
and high turbulence on the top and around buildings. Also, in
economic terms, the high costs of wind measurements campaigns
are an important barrier to the development of this sub-sector of
wind energy. Other data sources may be used for the characterization of the wind ow in urban environment, such as databases
and national and regional wind potential atlas. These solutions are
usually based on the application of data from mesoscale models
(MM51 e Fifth Generation Mesoscale Model, WRF e Weather

Abbreviations: MM5, Fifth Generation Mesoscale Model; WRF, Weather


Research and Forecasting; CFD, Computational Fluid Dynamics; U-DTM, Urban
Digital Terrain Model; PD, Power Density.
* Corresponding author.
~es), ana.estanqueiro@ineti.pt
E-mail addresses: teresa.simoes@ineti.pt (T. Simo
(A. Estanqueiro).
http://dx.doi.org/10.1016/j.renene.2015.12.008
0960-1481/ 2015 Elsevier Ltd. All rights reserved.

Research and Forecasting) to standard microscale models (e.g.


WASP - Wind Atlas Analysis and Application Program [2]) that
despite their validity, are not adapted to these environments. In
both methods, the wind potential is often overestimated [3].
The use of CFD (computational uid dynamic) models to characterize the wind behavior around buildings is nowadays state of
the art. Nevertheless, the application of these models is highly time
consuming, in particular when one needs to model large areas to
adequately assess the impact of structures on the wind ow. The
complexity of the domain's geometry requires the use of powerful
computers in order to obtain reliable results, and reinforces the
non-suitability of most CFD programs to the study of large areas. In
this context, a methodology was developed to characterize the
urban wind potential, based on the construction of a surface
involving the buildings' area, so that it can be treated as a very
complex orography. This methodology considers the application of
CFD models to small areas of a certain urban region, enabling the
establishment of correction factors to the overall spatial distribution of the wind ow. In recent years the CFD sector has developed
several commercial products, with some of them especially oriented to urban wind applications. In this particular case, the
WindSim model [4] (referred subsequently as CFD-Complex) is
used to model the surface of the buildings and the surrounding
terrain (urban digital terrain model: U-DTM) and the Meteodyn
model [5] (referred subsequently as CFD-Urban) is used to model

~es, A. Estanqueiro / Renewable Energy 89 (2016) 598e605


T. Simo

599

the natural geometry of the buildings in a small area. In both cases,


the wind data from the Wind Potential Atlas for Mainland Portugal
[6] are used. The results obtained by each model are compared and
discussed in this paper.

purpose of calibrating results with simple standard models for a


large city area. The models here used solve the RANS equations for
an uncompressible ow (Eqs (1) and (2)) and the k turbulence
model (Eqs. (3) and (4)) [17,18].

2. The wind in urban environment

"
!
#


v ruj ui
vp
v
vui vuj
0 0



m
 rui uj fi 0 ;
vxj vxj
vxj vxi
vxj

The urban wind resource has been object of several studies.


Although there are few references to urban wind energy planning
in a smart city context, there are a few R&D projects dedicated to
the study of the wind behavior around and above small groups of
buildings aiming at the installation of small scale energy systems,
e.g. Ref. [7].
One of the most important barriers for the study of the wind in
urban environments is the lack of adequate wind data measurements and CFD models are frequently used to ll this gap. Ultrasound and conventional anemometry is sometimes used to assess
the wind in urban areas, mainly in the scope of R&D projects [8].
LiDAR (Light Detection And Ranging) measurement systems [9] and
statistical methods such as the Weibull distribution based on large
wind databases have also been used by the scientic community for
this purpose, although the scale of application is, in most cases, of
the order of few tens of meters [10]. The physical representation of
the urban fabric is of outmost importance, independently of the
methodology applied for the wind measurement and its energy
resource assessment. . Most of the areas considered in these studies
are in the order of tens of meters with maximum dimensions of
approximately 2000 m. Although the methodologies developed are
suitable to neighborhood and street scales, they often do not apply
to city scale or larger areas [11]. The NUDAPT database project National Urban Database with Access Portal Tool, is an exception. It
contains 2D and 3D information on a set of American cities and
allows the user to download and use it in research projects. The use
of GIS e Geographic Information Systems e is also frequent to
model the urban fabric in larger areas, especially when these areas
are located in complex terrains [12].
2.1. Models for the urban wind resource assessment
Models based on the potential ow theory are widely used for
the development of wind parks but they do not usually have the
capacity to model the wind ow in urban areas (e.g. WASP). As an
alternative, CFD numerical models have been widely used in recent
years. These models have high computational requirements when
compared to standard models, but technology advances in the
latest years made their use in simulation possible in nearly standard computers especially when applied to small areas. The
application of CFD models for wind resource assessment is performed by solving the RANS equations e Reynolds Averaged
Navier-Stokes e with turbulence closure [13,14], where the turbulence models most commonly used are k and ku. CFD models
with mesoscale outputs as boundary and initial conditions, or with
time varying boundary conditions, have been applied by some
authors [15,16]. Their use is justied by the fact that the sole use of
mesoscale models for this purpose will overestimate the wind
energy obtained at the city scale.
Although CFD models are state of the art in urban wind
modeling, its application presents some disadvantages: they are
highly computation time-consuming; the construction of the
domain geometry is difcult and complex; and convergence
problems often occur introducing errors in the results. These disadvantages are important in the choice of the methodology to
follow when characterizing the wind resource, especially in
extensive areas. Nevertheless, in this work a CFD model for urban
environments (CFD-Urban) was used in small areas, with the main

vrui
0

vxi

(1)

(2)
0

where the term rui uj represents the Reynolds tension tensor, that
corresponds to the additional transference of momentum due to
the turbulence uctuations [19,20].



v
v nT vk
PK 
Ui k
vxi
vxi sk vxi

(3)



v
v nT v

2
C1 PK  C2
Ui
vxi
vxi s vxi
k
k

(4)

where Cm, sk, s, C1 and C2 are known constants, PK is the turbulent kinetic energy production given by Eq. (5), and nT is the kinematic viscosity.
The CFD-Complex model solves the RANS equations using the
nite element method, along with initial conditions, referred to as
the solution's rst estimates. This model may use two types of
solver e namely Segregated and Coupled. The Coupled solver
(Migal) [21] refers to the coupling technique speed e pressure and
to a linear solver that simultaneously updates the speed and
pressure leds across the domain. The linear solver performs only
the rst part of the speed-pressure iteration and is followed by,
Phoenics [22] model that completes the non-linear part of the
iteration. The CFD-Complex process involves the execution of
several modules which includes models that are executed
sequentially in order to obtain the results [18].
The CFD-Urban model also solves the RANS equations and uses
the nite volume method and a turbulence closure. The boundary
conditions and the domain are generated automatically and the
wind proles are obtained according to a theoretical reference
wind. The turbulent kinetic energy is constant and it is evaluated
according to the input roughness data from the simulation domain
[13]. Both models use wind frequency distributions (Weibull distributions and wind rose) as input data.
3. Methodology for the wind resource assessment in urban
environments
The methodology presented in this paper is based on the generation of a digital terrain model which includes the terrain and the
existing buildings as a whole, thus representing an urban digital
terrain model: U-DTM. The urban DTM can be treated as a very
complex terrain and be used as input for a standard wind resource
assessment model (e.g. Wasp, WindSim). This methodology
strongly reduces the computational costs associated with standard
CFD models to simulate groups of buildings; it simplies the geometry of the urban mesh and permits to extend the area of
simulation to a city scale.
The U-DTM is then inserted into the CFD-Complex and, by using
synthetic wind data series obtained by mesoscale modeling, which
enables the urban wind potential to be estimated in the absence of
experimental wind data. Since the U-DTM will, in some regions,
smooth the city geometry, and since wind data from numerical
mesoscale modeling usually overestimate the wind potential in

~es, A. Estanqueiro / Renewable Energy 89 (2016) 598e605


T. Simo

600

urban environments, it is also necessary to correct the spatial wind


resource maps obtained in order to describe the urban wind ow
more accurately. Therefore, in the proposed methodology some
areas of a city are selected and modeled in CFD-Urban in order to
establish correction factors to be applied to the results from the
CFD-Complex. The results from the two models are compared and
the deviations analyzed to dene the correction factors to be
applied to the wind resource maps obtained with the U-DTM. The
correction factors are obtained through Eq. (5).


Fci 1 

vref


(5)

vmodel i

where Fci is the correction factor, vre is the mean wind speed obtained with CFD-Urban and vmodeli is the mean wind speed obtained
with the CFD-Complex with the digital surface model for grid point
i. Depending on the extension of the urban area under analysis,
correction factors are dened for the area surrounding the buildings (area A) and for the building's rooftops (Area B), using the
algorithm described in (6).


vi

Fcib  vmodeli if Pi 2farea Bg


Fcig  vmodel i if Pi 2farea Ag

(6)

where Fcib is the correction factor obtained for buildings area, Fcig is
the correction factor for an area above the ground and without
buildings and Pi is the grid point i in the mean wind speed spatial
distribution map. For extended urban regions (e.g. area > 5 km2), it
may not be suitable to use the detail expressed in (7) due to the
slow pace and complexity that this procedure will bring to the
method. In this case a mean correction factor can be used for the
whole area (expression 7):

vi Fcim  vmodel i

(7)

where Fcim is the mean correction factor obtained from the comparison between the results obtained in both simulations. The same
formulation is performed for power density calculation (W/m2).
The correction factors are then applied to the resource map(s) obtained with the application of standard wind ow models to the UDTM in order to obtain a more accurate spatial description of the
wind resource.
4. Application of the methodology to an urban area
To test the proposed methodology, two neighbor areas were
~o Domingos
selected inside Cascais municipality: Estoril and Sa

administrative regions e where the estimation of the wind potential was performed using the models and approach described in
the previous section. Synthetic wind data series obtained by
mesoscale modeling were used as wind resource input for the
models [23]. In the CFD-Complex approach, the orography of the
city surroundings (10  10 m orography lines) was used together
with a superimposed surface generated from the 3D geometric
information about the buildings in that area, enabling the generation of an urban digital terrain model (4  4m orography lines) as in
Fig. 1. In CFD-Urban, the 3D detailed geometry of the building was
used to model small sets of buildings in the city under study, according to the methodology described in Section 3.
The data obtained from the Portuguese Wind Potential Atlas is
referred to grid points distributed on the areas under study. CFDComplex was used to model the wind speed over the U-DTM and
therefore to obtain wind resource maps. Fig. 2 represents the mean
wind speed and power density at h 10 m obtained with CFD~o Domingos villages, respectively.
Complex in Estoril and Sa
~o Domingos were then
Some regions inside Estoril and Sa
modeled with the natural geometry of the buildings, using the CFDUrban model to establish the correction factors in order to improve
the precision of the U-DTM methodology results.
~o Domingos
Fig. 3 presents the modeled areas of Estoril and Sa
used to establish correction factors in order to obtain the nal wind
resource maps for the urban areas under study.
Figs. 4e7 show the mean wind speed and power density spatial
distribution for the urban areas represented in Fig. 3 for h 10 m. In
order to establish correction factors for the results obtained with
the surface methodology, several locations were selected in the
areas under study and the mean wind speed and power density
results were compared.
~o
Figs. 8 and 9 show the chosen locations in Estoril and Sa
Domingos villages, respectively. Table 1 shows the deviations between the results obtained with both models.
As shown in Table 1, when considering the locations above the
buildings, the deviations are in the order of 2%. For the open nonbuilt areas the deviations are larger, the average being in the order of 13%. The mean deviation for the whole area is 6.65%. It should
be noticed that some of the locations present larger deviations than
others especially in open non-built areas. This is mainly due to the
fact that they correspond to densely urbanized zones, which are
less well described by the U-DTM and where skimming ow has a
strong chance to occur.
The same evaluation was performed for power density (FPI).
Table 2 shows the results obtained for both villages. In this case the
deviations are in the order of 1% for the building rooftops, 23% for

Fig. 1. (a) Extract of the urban area under study, (b) 3D representation of the generated Digital Terrain Model.

~es, A. Estanqueiro / Renewable Energy 89 (2016) 598e605


T. Simo

601

Fig. 2. Spatial distribution of the mean wind speed and power density in Estoril and Sao Domingos. h 10 m.

Fig. 3. Estoril (top) and Sao Domingos (bottom) areas modeled with CFD-Urban.

Fig. 4. Spatial distribution of the mean wind speed for h 10 m above rooftops in Estoril.

the building-free areas and 10% for the whole analyzed areas.
The nal mean wind speed and power density spatial distribution maps were obtained by the application of the mean correction
factors to the overall area under study, being 6.65% for the mean
wind speed and 10% for the power density. Fig. 10 depicts the nal
maps for the regions under study.
5. Validation and discussion of the results
In order to validate the results obtained with the U-DTM

methodology (CFD-Complex model), a LiDAR system was installed


on a building's rooftop located in Estoril village (data used is
referred to the period June 2012 to May 2013 with inter-annual
correction of the horizontal mean wind speed values). In addition, an anemometric station is operating at a small airport located
~o Domingos since 2004 - Tires, and a period between 2009 and
in Sa
2012 was used for the validation analysis. The validation analysis
was performed under three situations for each urban area. In Estoril
the measurements were compared with the CFD-Complex results
for mean wind speed and power density with the application of: i)

602

~es, A. Estanqueiro / Renewable Energy 89 (2016) 598e605


T. Simo

Fig. 5. Spatial distribution of the power density h 10 m above rooftops in Estoril.

Fig. 6. Spatial distribution of the mean wind speed for h 10 m above rooftops in Sao Domingos.

Fig. 7. Spatial distribution of the power density for h 10 m above rooftops in Sao Domingos.

rooftop correction factor e once the LiDAR is installed in the top of a


building: ii) mean correction factor; and iii) no correction at all. For
Sao Domingos; a) free building correction factor (for open and nonconstructed areas) was applied since the anemometric station is
located on the ground, b) mean correction factor and c) no
correction at all. Table 3 shows the deviations obtained between
the estimates and the measurements.
The deviations obtained with the application of the mean factor
are smaller in Estoril than in S~
ao Domingos, and in this last case the
application of the free building factor would approximate the estimates to the measured wind speed. The opposite happens when
the power density is under evaluation. This is mainly due to the

specicity of the urban area. The building geometry, height and


spatial density are important factors for the study of the wind
behavior, but it should be noted that one of the objectives of this
work is to develop a methodology that can be replicated to other
urban areas. Therefore the application of a mean factor should
prevail. Moreover, there are few measurement points, which inuence this analysis. It will be important to add more validation
points in the future, to strengthen the procedure. The vertical
prole was evaluated in Estoril case study since this was the only
experimental campaign where the use of a LiDAR station enabled
measurements at different heights. The wind speed vertical prole
is shown if Fig. 11 and presents a more pronounced development

~es, A. Estanqueiro / Renewable Energy 89 (2016) 598e605


T. Simo

603

Fig. 8. Selected test locations for Estoril.

~o Domingos.
Fig. 9. Selected test locations for Sa

Table 1
Comparison results for mean wind speed obtained with both models in Estoril and Sao Domingos (h 10 m).
ID

Estoril

Sao Domingos

Area 1

1
2
3
4
5
6
7
8
9
10
11

Area 2

Area 3

Area 1

Area 2

CFD-Urban

CFD-Complex

CFD-Urban

CFD-Complex

CFD-Urban

CFD-Complex

CFD-Urban

CFD-Complex

CFD-Urban

CFD-Complex

4.7
5.11
5.29
4.65
5.69
4.9
3.79
3.37
3.38
3.46
3.77

5.18
5.53
5.48
5.09
5.26
5.09
5.40
5.50
5.20
5.40
5.20

5.03
5.11
5.24
4.6
5.23
5.71
4.13
4.35
4.1
4.51
4.88

5.53
5.64
5.36
5.52
5.44
5.4
5.65
5.42
5.4
5.35
5.47

5.33
5.03
5.14
4.86
5.05
5.04
4.59
4.49
4.52
4.8
4.68

4.95
4.97
4.89
4.96
4.95
4.93
4.93
5.08
5.01
4.84
4.91

5.16
4.14
4.66
5.10
5.06
4.72
4.63
4.52
4.85
4.73
e

4.66
4.53
4.67
4.8
4.69
4.73
4.81
4.66
4.59
4.76
e

5.06
4.61
4.23
4.51
4.67
4.81
4.64
5.09
4.98
5.07
5.05

4.45
4.52
4.75
4.72
5.02
4.6
4.87
4.86
4.72
4.47
4.82

above h 20 m. Between 10 m and 20 m the development is slow


and is related to the existence of higher turbulence immediately
above the building rooftop. The estimates after the application of
the mean correction factor are close to the measurements, which
reinforce the validity of the developed methodology.
Although the validation process was performed with only two
measurement sites, the deviations obtained are acceptable and in
agreement with the local measurements, and also with the results
presented by other authors. As an example, Milashuk and Crane
[24] have developed a methodology based on CFD simulations that

was validated against 4 measurement sites and obtained deviations


in the order of 3%, which are similar to those obtained with the
methodology developed. Weekes and Tomlin [25], tested 3
different models, which they validated against 38 measurement
sites located in different terrain types. Considering the overall
terrain typologies, the deviations were in the order of 16% for mean
wind speed and 63% for power density. In the particular case of
urban environments, the obtained deviations were in the order of
18% for mean wind speed and 89% for power density, values
considerably higher than the obtained with the present

~es, A. Estanqueiro / Renewable Energy 89 (2016) 598e605


T. Simo

604

Table 2
Comparison results for power density obtained with both models in Estoril and Sao Domingos(h 10 m).
ID

Estoril

Sao Domingos

Area 1

1
2
3
4
5
6
7
8
9
10
11

Area 2

Area 3

Area 1

Area 2

CFD-Complex

CFD-Urban

CFD-Complex

CFD-Urban

CFD-Complex

CFD-Urban

CFD-Complex

CFD-Urban

CFD-Complex

CFD-Urban

102
115
149
98
184
123
53
39
43
44
55

141
169
163
159
145
139
161
165
150
156
145

131
136
149
100
149
190
73
84
73
95
120

167
178
151
165
158
156
178
157
157
150
159

158
129
137
118
132
126
101
91
90
115
106

119
123
115
118
117
118
116
128
124
113
118

133
71
103
131
134
143
133
129
124
148
e

100
91
99
109
102
104
109
99
94
107
e

127
92
74
120
97
106
65
96
85
92
93

84
88
101
99
121
93
109
111
101
87
110

Fig. 10. Spatial distribution of the mean wind speed and power density, h 10 m in Estoril and Sao Domingos after the application of the correction factors.

Table 3
Comparison between estimated mean wind speed values (V) and Power density (PD) with measurements for both villages.

a)Estoril
~o Domingos
a)Sa
b)Estoril
b)S~
ao Domingos
c)Estoril
c)S~
ao Domingos

Correction
factor for V [%]

V [m/s]
CFDcomplex
H 10 m

V [m/s] LIDAR
H 10 m

98
87
93.4
93.4
100
100

5.07
4.65
4.83
4.99
5.17
5.34

4.88
e
4.88
e
4.88
e

V [m/s] Tires
H 10 m

4.47
e
4.47
4.47

methodology, especially taking into consideration that both CFD


methods are very demanding in computational terms thus preventing their application to extensive urban areas.
The input data are obtained by numerical mesoscale simulation,
which constitutes a limitation of the method, since it does not
enable to represent, with the desired precision, the local effects
introduced by the presence of the city obstacles (in this case, the
buildings). This is noted in the differences observed in the Weibull
parameters (A and k) between the input data and the measurements, especially in the k parameter(overestimated in the mesoscale simulation between 25% and 33%), that being determinant for
the calculation of the power density. It is also important to refer the
selected simulation height. Abohela et al. [26] studied the height
above roof for urban wind resource assessment. Several heights are
mentioned in their work, varying between 1.3 h and 2 h according
to the relative height of the building and its surroundings. Other
authors have focused their research on this issue, the majority

Deviation
for V [%]

Correction factor
for FPI [%]

PD [W/m2]
CFD- complex
H 10 m

PD [W/m2]
LIDAR
H 10 m

PD [W/m2]
Tires
H 10 m

Deviation
for PD [%]

3.89
4.08
 1.02
10.6
5.94
19.7

101
77
90
90
100
100

142
102
127
120
141
133

161
e
161
e
161
e

e
123
e
123
e
123

11.8
17.07
21.1
2.43
12.4
8.13

referring the interval 1 he2 h as suitable for small wind turbines


support tower installation. In the present work, the main objective
was to develop a general methodology that can be replicated to
several and extensive urban areas. The introduction of a factor such
as the building height variation would make replication difcult
due to the added complexity that it would bring to the method. The
selection of a standard xed height will allow the replication of the
methodology since the resource maps can be developed for any
height above the generated U-DTM enabling the identication of
suitable areas for urban wind resource exploitation for any height
above the rooftops.
6. Conclusions
This paper presented a methodology for urban wind resource
assessment based on the generation of a surface involving the urban fabric and surrounding terrain. The proposed methodology is

~es, A. Estanqueiro / Renewable Energy 89 (2016) 598e605


T. Simo

Fig. 11. Vertical mean wind speed proles estimated by the CFD-ORO model (with and
without mean correction factor) and measured by the LIDAR.

based on simple and low-cost procedures that allow its replication


to other urban areas. The results obtained were validated against
measurement data and the estimates were within acceptable deviations with respect to wind data point measurements. The use of
the U-DTM methodology proved to be a valuable approach to
obtain the sustainable wind potential in urban areas of large dimensions (e.g. on a city scale) thus enabling the identication of
suitable areas for the installation of small wind turbines. The generation of the urban digital terrain model was based on well-known
methodologies and can be performed by less-experienced users on
a daily basis. The results obtained are a valuable contribution for
the calculation of the sustainable wind resource in urban areas, for
planning purposes, especially in the context of smart cities.
Acknowledgments
The authors wish to thank the municipality of Cascais and Prof.
nio Lopes from IGOT - Institute of Geography and Spatial
Anto
Planning of the University of Lisbon for providing the 3D-cartography of the municipality, including the buildings geometry, and
the access to wind data measurements from the Tires airport. The
authors also would like to thank the municipality of Cascais for
allowing the installation of a LIDAR system at Pedra do Sal and to
Dr. Margarida Fontes from LNEG for the valuable help in the text
review of this paper. The work developed for this paper was funded
by the Portuguese Foundation for Science and Technology (FCT)
under the grant: SFRH/BD/41756/2007 and by the National Laboratory for Energy and Geology, I.P. (LNEG I.P.).
References
[1] Sinisa Stankovich, Neil Campbell, Alan Harries, Urban Wind Energy.: Earthscan, BDSP Partnership Ltd, 2009.

605

[2] N.G. Mortensen, L. Landberg, I. Troen, E.L. Petersen, Wind Atlas Analysis and
Application Program (WASP), vol. 1, 1983. Getting started, Roskilde.
~ es, P. Costa, A. Estanqueiro, A rst methodology for wind energy
[3] T. Simo
resource assessment in urbanised areas in Portugal, in: 2009 European Wind
Energy Conference Proceedings, Marseille, 2009.
[4] WindSim, WindSim, Version 4.9.1.24234, copyright 1997-2009, WindSim
AS, 2009.
[5] Meteodyn, UrbaWind, Version 1.5.0.0, copyright , Meteodyn, 2008, p. 2008.
[6] P. Costa, P. Miranda, A. Estanqueiro, Development and validation of the Portuguese Wind Atlas, in: 2006 European Wind Energy Conference Procedings,
Athens, 2006.
[7] Dong Li, Shujie Wang, Peng Yuan, A Review of micro wind turbines in the built
environment, in: Power and Energy Engineering Conference (APPEEC), 2010
Asia-Pacic
1,
2010,
pp.
1e4,
http://dx.doi.org/10.1109/
APPEEC.2010.5448223, 4, 28e31 March 2010.
[8] K. Syngellakis, P. Clement, J. Cace, Wind Energy Integration in the Urban
Environment - WINEUR, 2005. U.K.
[9] S. Gryning, E. Batchvarova, Measuring meteorology in urban areas e some
progress and many problems, in: Meteorological and Air Quality Models,
Springer, Berlin, 2009, pp. 125e132.
[10] J.L. Acosta, S.Z. Djokic, Assessment of renewable wind resources in UK urban
areas, in: MELECON 2010-2010 15th IEEE Mediterranean Electrotechnical
Conference, 2010, pp. 1439e1444 [Online], http://ieeexplore.ieee.org/xpls/
abs_all.jsp?arnumber5476217&tag1.
[11] R. Britter, T. Bentham, Spatially averaged ow within obstacle arrays, Atmos.
Environ. 37 (2003) 2037e2043.
[12] M. Kanda, Review of Japanese urban models and a scale model experiment, in:
Meteorological and Air Quality Models for Urban Areas, Springer, Berlin, 2009,
pp. 39e46.
[13] A. Kalmikov, G. Dupont, K. Dykes, C. Chan, Wind power resource assessment
in complex urban environments. MIT campus case-study using CFD analysis,
in: AWEA 2010 Windpower Conference, May 23-26, 2010, pp. 1e28.
[14] L. Ledo, P.B. Kosasih, P. Cooper, Roof mounting site analysis for micro-wind
turbines, Renew. Energy 36 (2011) 1379e1391.
[15] M. Tewari, et al., Impact of coupling a microscale computational uid dynamics model with a mesoscale model on urban scale contaminant transport
and dispersion, Atmos. Res. 96 (2010) 656e664.
[16] J. Baik, J. Kim, Urban ow and dispersion simulation using a CFD model
coupled to mesoscale model, Am. Meteorological Soc. (2009) 1667e1681.
[17] Tristan Wallbank, WindSim Validation Study - CFD Validation in Complex
Terrain, WindSim AS, Tnsberg, 2008 [Online], http://www.windsim.com/
documentation/papers_presentations/thesis/080512trw%20WindSim%
20Write%20Up%20-%20Validation%20study.pdf.
[18] Arne R. Gravdahl, Meso Scale Modeling with a Reynolds Averaged NavierStokes Solver. Assessment of wind resources along the Norwegian coast, in:
31th IEA Experts Meeting, State of the Art on Wind Resource Estimation,
RISOE, 1998. Roskilde.
[19] N.O. Jensen, N.E. Bush, Atmospheric turbulence, in: Engineering Meteorology,
Elsevier, Amsterdam, 1982, pp. 179e231.
[20] J.R. Holton, An Introduction to Dynamic Meteorology, third ed., Elsevier Science (USA), San Diego, USA, 1972.
[21] Michel Ferry, New features of the MIGAL solver, in: 9th Phoenics User Conference, 2002, pp. 1e24. Moscow, http://www.cham.co.uk/PUC/PUC_Moscow/
MRFDC.pdf [Online].
[22] CHAM Ltd, PHEONICS - Your Gateway to Successful CFD, 2013 [Online], http://
www.cham.co.uk/.
[23] P. Costa, P. Miranda, A. Estanqueiro, Development and validation of the Portuguese wind Atlas, in: EWEC 2006-European Wind Energy Conference 2006,
2006, pp. 1e9. Athens.
[24] S. Milashuk, W.A. Crane, Wind speed prediction accuracy and expected errors
of RANS equations in low relief inland terrain for wind resource assessment
purposes, Environ. Model. Softw. 26 (4) (Apr. 2011) 429e433 [Online], http://
www.sciencedirect.com/science/article/pii/S136481521000263X.
[25] S.M. Weekes, A.S. Tomlin, Evaluation of a semi-empirical model for predicting
the wind energy resource relevant to small-scale wind turbines, Renew. Energy 50 (2013) 280e288.
[26] Islam Abohela, Neveen Hamza, Steven Dudek, Effect of roof shape, wind
direction, building height and urban conguration on the energy yeld and
positioning of roof mounted wind turbines, Renew. Energy 50 (Feb. 2013)
1106e1118
[Online],
http://www.sciencedirect.com/science/article/pii/
S0960148112005381.

You might also like