Microeolic Turbines in The Built Environment in Uence of The Installation Site On The Potential Energy Yield

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Renewable Energy 45 (2012) 163e174

Contents lists available at SciVerse ScienceDirect

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

Microeolic turbines in the built environment: Influence of the installation site


on the potential energy yield
Francesco Balduzzi, Alessandro Bianchini, Lorenzo Ferrari*
“Sergio Stecco” Department of Energy Engineering, University of Florence, Via di Santa Marta 3, 50139 Florence, Italy

a r t i c l e i n f o a b s t r a c t

Article history: Generic proposals for an effective integration of renewable energy sources in the urban environment are
Received 27 September 2011 frequently carried out by project developers, local governments and media, although an in-depth
Accepted 25 February 2012 knowledge of the technical and energetic limitations is often missing. In particular, the installation of
Available online 23 March 2012
small wind turbines on the rooftops of tall buildings is considered to represent an attractive solution
thanks to the supposed possibility of exploiting local flow accelerations induced by the building façades.
Keywords:
The real feasibility of this scenario has, however, yet to be proved, both in terms of real energy harvesting
Urban environment
and of compatibility of the machines with a densely populated area.
Microeolic
Wind turbine
In this study, a critical examination of the flow conditions on the rooftop of a building in an urban
Rooftop environment has been carried out by means of CFD simulations. The main goal of the analysis was the
CFD assessment of some general criteria to evaluate the convenience of a microeolic turbine installation on
the roof of a selected building as a function of both its geometrical features and those of its upwind
building along the prevailing wind direction. In each configuration, the flow velocity and skew angle on
the rooftop area of the installation building were calculated for different incoming wind profiles and
compared to their levels in the undisturbed flow. Finally, an energy-based comparison between the flow
potentials in the investigated environment is provided and some general tendencies are outlined.
Ó 2012 Elsevier Ltd. All rights reserved.

1. Introduction reduction of the capital cost of the machine and of the instal-
lation cost, as well.
Increasing efforts are being made in order to reduce the ener- - Delocalized power production: a high number of installations
getic demand of urban centers; in particular, small-size renewable could be achieved. In addition, the required power would be
energy systems are thought to represent a very interesting solution generated on-site, making the dispatch system simpler and
for achieving a form of sustainable and low-energy design in new more efficient.
buildings. Among others, the installation of small and medium-size - Integration of renewable energy systems: the introduction of
wind turbines on the rooftop area of high buildings has been often renewables in the urban environment could provide relevant
suggested by architects and project developers as a valuable solu- benefits in terms of both pollution reduction and fuel saving.
tion (e.g. Refs. [1e5]). The clear advantages of a similar installation
context can be readily understood: Despite these interesting remarks, the wind conditions in the
rooftop areas of buildings in urban locations are indeed very
- Higher installations: small rotors positioned at the top of a tall complex and the real adaptability of wind turbines to this context
building could reach installation heights that would not be has not yet been assessed, both in terms of real producibility and of
allowed by the only turbine tower, then exploiting a faster flow compatibility with the environment itself. In further detail, the
in this portion of the vertical wind profile; at the same time, feasibility of this type of installations in terms of energy production
high towers would not be needed anymore, with a notable has not been verified on the basis of some perplexities that are still
connected to the influence on the available wind velocity of the
complex geometry of the fabric of the city. Moreover, even if
* Corresponding author. Tel.: þ39 055 4796570; fax: þ39 055 4796342.
a suitable energetic potential is ensured, additional doubts arise
E-mail addresses: balduzzi@vega.de.unifi.it (F. Balduzzi), bianchini@
vega.de.unifi.it (A. Bianchini), ferrari@vega.de.unifi.it (L. Ferrari). from the visual and acoustic impact of the rotors and their struc-
URL: http://vega.de.unifi.it tural safety in a populated zone, as well.

0960-1481/$ e see front matter Ó 2012 Elsevier Ltd. All rights reserved.
doi:10.1016/j.renene.2012.02.022
164 F. Balduzzi et al. / Renewable Energy 45 (2012) 163e174

Focusing on the installation requirements, some general prin- installation building is often chosen to ensure a façade perpen-
ciples for identifying the main characteristics of a suitable site in dicular to the prevailing wind.
the built environment have been provided in the technical litera- On the basis of this scheme, 2-D simulations were carried out on
ture (e.g. Refs. [6,7]); in a first approximation, however, rooftop- different buildings configurations by assuming the turbine installed
installed wind turbines require an installation building reason- near the edge of the façade perpendicular to the prevailing wind
ably higher than the average height of the surrounding buildings direction. A realistic urban wind profile (in terms of both velocity
[1,7,8]. modulus and displacement) was then re-created along this direc-
The wind profile in urban locations is in fact quite different from tion in front of the installation building by introducing three solid
the classical log-law based profile [1,9], with the zero velocity elements with a constant height equal to the mean buildings’
height shifted up to a non-zero value (displacement, d) which is height of the investigated city.
a function of the mean height of the surrounding buildings (see In detail, the roughness parameters and the wind characteristics
Fig. 1). exploited in the simulations were extrapolated from a real city data
In these conditions, wind turbines which can be positioned on set; literature data for London city (Refs. [15,16]) were chosen
high buildings can indeed take advantage of the local flow devia- (Table 1). The displacement value was calculated by Eq. (1),
tion since they operate in a wind accelerated by the building itself following the indications of the ESDU [17]:
[1,7], but only on the condition that specific geometric proportions
between the installation building (IB) and its surrounding buildings ^  4:3z ð1  A Þ ¼ 13 m
d ¼ H (1)
0 h
are met [7]. Focusing the attention, however, on the evaluation
criteria of a specific installation site, general purpose schemes are
usually missing, due to the huge number of variables required for 3. Numerical analysis
an accurate analysis of the flow structure in the rooftop area. As
a consequence, the numerical investigations available are generally The CFD simulations were carried out with the OpenFOAM
focused on a well-defined urban environment or a specific case software package [18]. The modeling approach was a 2-D steady-
study (e.g. Refs. [10,11]). state Reynolds-Averaged NaviereStokes (RANS) calculation for
incompressible flow without buoyancy effects. From a fluid-
2. Case study and theoretical approach dynamic point of view, the assumption of a 2-D behavior of the
flux that invests the turbine implies that the streamlines are
The flow field in the rooftop of a building is very complex and assumed to follow the same path in each plane perpendicular to the
fully three-dimensional; in addition, the velocity modulus and building façade. This reasonable hypothesis, often applied in
directions are strongly influenced by the geometry of the building numerical studies of the built environment (e.g. Refs. [7,19e21])
itself and its surroundings, as well as by all the obstacles that are remarkably simplified the numerical model, providing a sensible
overcome by the wind along its approaching path toward the reduction of the computational cost as well as a finer in-plane
building. Based on the above, general tendencies to estimate the discretization.
convenience of a wind turbine installation in similar sites are quite
difficult to define and some substantial approximations are 3.1. Code settings
required.
Two crucial factors which must be accounted for, however, are The pressureevelocity coupling was made with the SIMPLE
the turbine layout and its integration with the building; in fact, if algorithm and the convergence to the final steady-state was
one neglects some futuristic solutions which integrate the rotor assessed with a target final computed normalized root-mean-
with the structure (e.g. Refs. [12,13]), the most suitable proposal for square residuals for all PDE’s of 1 105 (Refs. [19e21]).
a wind turbine use in the built environment is the installation of the The simulations were based on a ke3 two-equation modeling of
rotor on the edge (or as close as possible to it) of the building’s roof, the turbulence. As confirmed by a survey of the technical literature,
in order to exploit the wall effect of the oncoming flow on the this approach is indeed widely used to computationally model the
façade (Refs. [1,14]). atmospheric boundary layer (see Refs. [19e25]).
In this configuration, the available wind rose near the rotor is The wall boundary conditions at the bottom of the computa-
strongly distorted with respect to that of the same site in an tional domain were based on the standard wall functions [26] with
undisturbed location [8]; the roof shape and extension can indeed the sand-grain based roughness modification [27]. The roughness
substantially modify the wind distributions along a wide range of effect was taken into account by imposing suitable values of the
incoming wind angles. As a result, the turbine installation is almost sand-grain roughness Ks and the roughness constant CS (CS ¼ 0O1)
ever related to the prevailing wind direction of the site and the which can be determined from the roughness length z0 by the Eq. (2)
(see Ref. [22])

z0
Ks ¼ 9:793 (2)
CS
Within this modeling, the near-wall cell size is determined by
imposing the condition that its first nodal point has a distance yP
from the wall which satisfies Eq. (3)

Table 1
London data.
ˇ
Mean building height (H) 13.6 m
Percentage of the total area occupied by buildings 55%
Roughness length (z0) 0.29 m
Friction velocity (u*) 0.55 m/s
Fig. 1. Wind profile visualization in the internal boundary layer of a built environment.
F. Balduzzi et al. / Renewable Energy 45 (2012) 163e174 165

yP >Ks (3) of the last element with the same displacement of the wind profile
imposed at the inlet boundary.
In further detail, it is worth pointing out that other two-
As suggested by Refs. [20,21], the dimensions of the computa-
equation models, like the keu and particularly the SST model
tional domain (see Fig. 3) depend on the height of the tallest
originally proposed by Menter [28,29], are deemed to provide
building in the domain (i.e. the installation building). In our study,
a significant improvement of accuracy over the ke3 models in many
H was selected as an analysis parameter, assuming three different
applications (e.g. external flows over the airfoils). The SST model
values depending on the case study. Within this approach, in order
can in fact make use of a direct application of the no-slip condition,
to avoid any influence of the domain size on the final solution, all
with no requirement of the wall functions, then incorporating the
runs were simulated with fixed dimensions of boundaries: being Z
roughness effects in the u wall boundary condition [28e30]. In the
the height of the tallest building tested in all the simulations (i.e.
present case, however, the application of a similar model is prac-
the maximum value assumed by H), the computational domain was
tically precluded by the high roughness (more than one order of
therefore given by extending its boundaries from the installation
magnitude higher to that conventionally assumed as the limit of
building center to 16Z vertically, 5Z upstream and 15Z downstream.
application of modified u wall boundary condition [31,32]) and the
The inlet boundary conditions were imposed by the assumption,
excessive mesh resolution hypothetically required [31]. Moreover,
proposed by Ref. [16], of a constant shear stress with the height (the
recent works (e.g. Ref. [33]) showed that even the keu models with
hypothesis has been also validated by Refs. [16,21,23,27]). By the
the wall functions are assumed to provide poorer performance with
assignment of the friction velocity u*, the resulting inlet velocity
the current OpenFOAM version than that obtainable by means of
profile was therefore a logarithmic boundary layer; k and 3 were
a ke3 model.
imposed by a constant and a hyperbolic function, respectively. The
In order to assess the best numerical configuration, the perfor-
final model settings are summarized in Table 2.
mance of several turbulence models was investigated and validated
The mesh discretization was obtained by a structured quadri-
with experimental wind tunnel measurements of the CEDVAL
lateral grid with high resolution: the expansion ratio of adjacent
laboratory [34] by means of a test case on a single building block
cells is kept below 1.15 (Refs. [19e21] suggested a value lower than
(1:200 scale). In particular the standard ke3 model, the RNG ke3
1.3).
model and the realizable ke3 model were compared: the standard
This setting was determined on the basis of a sensitivity analysis
ke3 model showed a better fit (see Fig. 2) with the experimental
on the cells’ number, in order to ensure the grid-independency of
data and it was therefore chosen as the base scheme for the whole
the results. The variation of the total amount of the cells was carried
set of simulations.
out by progressively reducing the maximum expansion ratio
The main problem in applying the scale factor to the final model
between adjacent cells from the limit value suggested by literature
was, however, to satisfy the yP constraint both at the ground and in
(1.3 in Refs. [19e21]) to a minimum of 1.08.
the building walls; this constraint would lead, indeed, for z0 values
In order to increase the level of confidence, the analysis was
in the urban context (w1.0 m) to the creation of excessively
repeated on two geometries included in the test plan, having
ˇ ˇ ˇ
extended cells at the ground (yP w 30 m) with respect to the near-
constant values of h ¼ 2H , D ¼ H and L ¼ H , whereas the IB height
ˇ
building size (z0 ¼ 3.0  104 m, yP w 0.01 m).
was varied from the minimum tested value (2H e Model 1) to the
ˇ
To overcome this criticality, an explicit modeling of the rough-
maximum one (4H e Model 2). In particular, five refinement levels
ness elements was applied: these elements were considered as
of the mesh were tested on both the configurations: the boundary
square blocks having the function to virtually reproduce the real
conditions and simulations settings shown in Table 2 were applied.
roughness effects on the flow. By doing so the ground roughness of
The results were analyzed on the basis of the variation in the
an open-landscape environment can be modeled using a smaller
calculated mean velocity in two regions of interest of the domain.
value of z0 (w3.0  102 m).
The first region is that 10 m above the upwind corner of IB, which
The roughness elements height (HR) and spacing were imposed
encloses the macroscopic flow structures of interest, i.e. the
ˇ
to be equal to the values of the average building height (H ) and the
streamlines curvature due to the overcoming of the building. Then,
percentage of the total area occupied by buildings of the investi-
the attention was focused on the 2 m height zone above the upwind
gated urban area, respectively. The number of the roughness
corner of IB, which was considered a suitable zone for a wind
elements was chosen in order to recreate a wind profile at the end
turbine functioning. The results of the mesh refinement in Model 2
(i.e. the most critical condition) are shown in Fig. 4. Upon exami-
nation of the trends of Fig. 4, a number of cells of about 8  104 was
selected as the best compromise between accuracy and

Fig. 2. Test case with the Standard ke3 turbulence model: comparison between
simulated and experimental vertical velocity profiles at different longitudinal
positions. Fig. 3. Analysis domain.
166 F. Balduzzi et al. / Renewable Energy 45 (2012) 163e174

Table 2 Table 3
Numerical model settings. Test plan.

Algorithm SIMPLE IB height (H) 2, 3, 4


Interpolation Schemes Always second order UB height (h) 1.00, 1.25, 1.50, 1.75, 2.00,
Turbulence Model Standard ke3 2.25, 2.50, 2.75, 3.00, 3.25
u* ðz  dÞ u* 3 u* 2 Distance between IB and UB (D) 0.5, 1, 2, 4
Boundary conditions Inlet uðzÞ ¼ ln ; 3 ðzÞ ¼ ; kðzÞ ¼ pffiffiffiffiffiffi
k z0 kz Cm UB width (L) 0.5, 1, 1.5, 2
Top Symmetry Plane
Outlet Zero gradient
Roughness length Ground 3.0  102 m
Wall 3.0  104 m on the average wind speed in the rooftop of IB, with a decrease in
the first case and an increase in the second one. Each configuration
(in the continuation of the study named Case 1 or Case 2 families of
computational efforts: the velocity error between the chosen test models, see Table 4) has been simulated with three roof
refinement level and the maximum one is lower than 0.1% in the typologies (indicated by letters from a to c): in detail, a flat roof (e.g.
10 m zone and 0.3% in the 2 m zone. tall modern buildings, skyscrapers, etc.) was compared with
a sloping roof, with two inclination angles of 8 (equal to the roof
3.2. Test plan slope in central Italy) and 18 (equal to the average slope of the
Italian roofs).
The aim of the analysis was to investigate the modifications in
terms of flow velocity occurring at the rooftop of IB when different 4. Results
H, h, D and L values are considered. The flow variations were
evaluated in terms of velocity modulus and skew angle (g) The numerical results of CFD simulations were analyzed and
approaching the turbine, i.e. the angle between the local velocity compared by means of the average velocity in a 2 m height zone
vector and the wind main direction (perpendicular to the façade). (thanks to the 2-D approximation) over the upwind corner of IB:
In further detail, a parametric analysis was conceived, in which this area was in fact supposed to be a suitable zone for the func-
the values of the four investigated parameters were defined ˇ
tioning of a roof-mounted wind turbine. The velocity variation
(Table 3) as a function of the mean buildings’ height H in the DU [%], with respect to the undisturbed wind profile at the same
considered urban environment; as a consequence, the numerical height and the skew angle g were therefore calculated on the
results are here presented in a dimensionless form, i.e. as a function
ˇ ˇ
hypothesis that the modification of the oncoming flow field due to
of H (e.g. H h H/H ). the turbine can be neglected. In the averaging process, a maximum
The minimum height that has been taken into account for the relative standard deviation (RSD), i.e. the mean discrepancy
installation building (IB) was at least twice the average city height, between a calculated value at a specific height within the 2 m zone
in order to ensure a suitable incident flow on the rooftop. With this and the average quantity considered in the analysis, of 3% on the
goal in mind, a detailed investigation was then undertaken on the velocity modulus and of 7% on the skew angle were noticed. This
influence of the upwind building (UB) height, in order to assess the particular approach to the data analysis ensures a direct compar-
h/H ratio which ensures the best flow conditions; depending on the ison of the investigated geometries in terms of their effect on the
IB height, the range between the average city height (i.e. h ¼ 1) and flow structure modification on the rooftop of the building; it is
the IB height was considered. worth pointing out, however, that the absolute energy content of
Characteristics distance values (D) have been selected to the flow itself in each configuration is function of the height of the
represent the real building spacing due to a street, a parking space specific building considered, as a taller building will allow the
or a square, while the UB width was varied in order to evaluate the hypothetic turbine to operate in a higher zone of the vertical wind
effects of the UB aspect ratio and the rooftop extension. profile, i.e. in a faster flow.
In addition to the parametric test plan of Table 3, a further set of Within the present analysis, the influence of turbulence effects
simulations has been carried out to investigate the influence of the on the functioning of the wind turbine was neglected and it may
IB roof shape on the oncoming flow. For this purpose, two specific require further evaluations in the near future. Temperature effects
geometric configurations in terms of IB height have been consid-
ˇ ˇ
on the flow were also excluded from the simulations.
ered, with H ¼ 2H and H ¼ 4H respectively: after a preliminary
analysis, these configurations had in fact showed an opposite effect
4.1. Buildings height

Upon examination of Fig. 5, where DU and g are reported as


a function of the upwind building height h for different H values (D
is fixed and equal to 0.5), some noteworthy variation trends can be
readily outlined. Focusing on Fig. 5a, if one analyzes a single curve

Table 4
Roof configurations.

IB height UB height Distance between UB width Inclination


(H) (h) IB and UB (D) (L) angle (c) ( )
Case 1a 2 1.50 0.5 1 0
Case 1b 8
Case 1c 18

Case 2a 4 2.75 0.5 1 0


Case 2b 8
Case 2c 18
Fig. 4. Sensitivity analysis on the mesh discretization e Model 2.
F. Balduzzi et al. / Renewable Energy 45 (2012) 163e174 167

reported, which actually show a clearly defined trend outlining an


increase of the optimal hopt value for higher H.
Moreover, it is also worth noticing that DU is always negative for
low installation building heights, whereas some velocity incre-
ments can be obtained whenever IB height is sufficiently higher
than the mean height of the surrounding buildings. This latter trend
can be appreciated in detail in Fig. 6a, where the velocity fields for
the investigated IB heights are compared: the tallest is IB, the
highest is the acceleration in its upwind corner. In addition, in
Fig. 6b the influence of the UB height variation on the flow field is
reported: the acceleration is maximized when UB is able to pre-
accelerate the flow toward the IB corner with the creation of
a sort of “aerodynamic ramp” to the corner itself.
Upon examination of the skew angle trends (Fig. 5b), one can
readily appreciate the dependence on the buildings shape. In
detail, an increase of H leads always to an increase of g, and an
increase of h leads to a g reduction. In conclusion the flow direc-
tion is directly controlled by the geometric proportion between UB
and IB.

4.2. Buildings distance

Focusing the attention on the most attractive solution (i.e.


H ¼ 4), the velocity variations are presented in Fig. 7a as a function
of the UB height for different distance values between the build-
ings; in addition, the geometric angles a and the skew angles g are
also reported (Fig. 7b and c, respectively), being a the angle
between the upwind corners of UB and IB (see Fig. 2).
Fig. 7a allows one to make some relevant markups. It is worth
pointing out that, when H is fixed up and quite higher than the
mean height of the surrounding buildings:

- No influence on the flow velocity is introduced by UB when h is


Fig. 5. Velocity variation (a) and skew angle (b) as a function of the upwind building
height (h) for different installation building heights (H) at D ¼ 0.5.
short (i.e. the disturbances do not reach the rooftop of IB);
- Velocity increments (i.e. positive DU) are attainable when D is
short;
(i.e. H, D and L ¼ const.), a characteristic trend of dependence of the - The increments become more relevant when event UB is taller
flow acceleration from the UB height can be noticed; in particular, than the average height of the surrounding buildings. The h
the maximum velocity increment is achieved by an optimal value of value which maximizes the flow velocity for each configura-
the upwind building height (hopt), which does not in fact corre- tion increases by decreasing the distance between the build-
spond to the minimum height. On the other hand, the influence of ings. Further increments of h from the optimal value make the
the IB height can be evaluated by a wider analysis of three curves velocity variation become sensibly negative;

Fig. 6. Velocity fields for: (a) different H with constant h ¼ 1.5 and D ¼ 0.5; (b) different h with constant H ¼ 4 and D ¼ 0.5.
168 F. Balduzzi et al. / Renewable Energy 45 (2012) 163e174

Fig. 8. Velocity fields for: (a) H ¼ 4  h ¼ 2.75  D ¼ 1 (b) H ¼ 4  h ¼ 2.75  D ¼ 4.

IB front wall, while the UB height doesn’t introduce any disturbance


at the IB rooftop level.
Based on the evidence provided by the analysis of Fig. 7, in Fig. 9
is presented a comparison between the geometrical configurations
which maximize DU for the IB heights of 3 and 4, respectively: the
optimal geometries are plotted as a function of the ratio between
the UB and IB heights (h/H) and the distance D. As one may notice,
a surprising matching between the relative trends was found; in
detail, for a fixed H, when D increases, the value of the upwind
building height (h) which maximizes DU decreases linearly with
the distance between the buildings. Conversely, when the two
buildings are close to each other (i.e. D is short), h must not be much
shorter than H in order to exploit the acceleration on the rooftop. In
broader terms, for a fixed D, a variation of H should be followed by
an equal variation of h (i.e. constant ratio h/H).

4.3. Upwind building width

Upon examination of the results for a fixed installation building


height (H), a relationship between the buildings heights and their
distance has been highlighted, suggesting that the flow modifica-
tion on the rooftop of IB are primarily affected by the angle (a)
between the upwind building roof and the linear path which
connects the upwind corners of UB and IB.

Fig. 7. Velocity variations (a), geometric angles a (b) and skew angles g (c) as a function of
the upwind building height (h) for different distances (D) between the buildings at H ¼ 4.

- The h values which maximize DU seem to be connected with


the deriving geometric a angle (see Fig. 8a). The best positive
DU values (gray colored points) have been found when an
a value between 30 and 40 is found (i.e. the gray dashed line).
Otherwise, whenever a < 25 , negative DU values were found
(Fig. 8b).

Fig. 7c confirms the interpretation suggested by the analysis of


Fig. 5b: either an increase of h or D leads to a monotonic decrease of
the skew angle g. The trend is analogous to that of the geometric
angle a, even if at short h values the skew angle stabilizes itself to
an almost constant value, being independent on the distance D. The Fig. 9. Optimal ratio between the buildings heights as a function of the distance
final value is in fact only affected by the flow turning caused by the between the buildings themselves.
F. Balduzzi et al. / Renewable Energy 45 (2012) 163e174 169

Fig. 11. Velocity variation (a) and skew angle g (b) as a function of the upwind building
width (L) for different global distances Q ¼ L þ D (H ¼ 4  h ¼ 2.75).

Fig. 10. (a) Optimal ratio between the buildings heights and a angles (b) as a function
shows that the extent of the velocity reduction is poor (approxi-
of the global distance Q ¼ L þ D between the upwind corners of the buildings.
mately 1% by quadruplicating L), which indicates that a comple-
mentary variation of L and D has a lower influence than that induced
As a result, the horizontal distance Q ¼ L þ D between the by a variation of their sum Q ¼ L þ D; on this basis, the gap between
upwind corners of UB and IB is proposed as the most relevant the curves at different Q values is remarkable (a doubled L þ D value
analysis parameters in evaluating the influence of the buildings leads to a 6% reduction of the velocity in the rooftop area).
distance.
The importance of the Q parameter can be readily appreciated in
4.4. Roof shape
Fig. 11, where the configurations which maximize DU for different L
values are reported: the points are plotted as a function of the
The main results of the analysis on the effects of the roof shape
global distance Q and the h/H ratio (Fig. 10a). The linear trend
are summarized in Fig. 13. Although this geometric feature of the
already noticed in Fig. 9 is here confirmed as well as the influence of
the a angle on the flow modifications on the rooftop of IB (the
optimal a values are reported in Fig. 10b); in addition, the Q
parameter made all the investigated geometries collapse on the
same trend. As a consequence, it is also worth noticing that, when
the distance L þ D decreases, the optimal angle between the
buildings corners must increase as the arctangent of the height
difference between IB and UB (i.e. the difference H  h).
Furthermore, the coupled effect of the UB width (L) and global
distance (Q ¼ L þ D), for a fixed buildings’ geometry
(H ¼ 4  h ¼ 2.75), has been examined in Fig. 11. Focusing the
attention on a single curve (i.e. Q ¼ const.), one can notice that an
increase of the UB width, being a unaltered, implies a reduction of
the velocity variation (Fig. 11a). This particular behavior can be
better explained upon examination of Fig. 12: a wider region of
reattaching flow over the UB roof, when L is long (Fig. 14b), in fact
reduces in fact the acceleration effect toward IB. The increase of the
attached-flow region is confirmed by Fig. 11b: the skew angle g
increases with L, due to a flow upwards turning closer to IB. On the
other hand, an increase of the global distance Q leads to a g Fig. 12. Velocity fields for: (a) H ¼ 4  h ¼ 2.75  D ¼ 2.0  L ¼ 0.5 (b)
reduction, as already mentioned in the Section 4.2. Finally, Fig. 11a H ¼ 4  h ¼ 2.75  D ¼ 0.5  L ¼ 2.0.
170 F. Balduzzi et al. / Renewable Energy 45 (2012) 163e174

oncoming wind speed, a little variation can be noticed in the low-


speeds range due to the increased reduction of the characteristic
Reynolds number at these velocities, which influences the
viscosity-driven features of the overall flow field; for the same
reason, this effect is deemed to be more pronounced in buildings
with small dimensions.
On this basis, a correct evaluation of the energy yield of a roof-
mounted wind turbine must take into account the proper estima-
tion of the power coefficient variation as a function of the wind
speed. In this study, however, thanks to the evidence provided by
Fig. 13, the velocity variation and the skew angle were assumed to
be independent from the oncoming wind.

5. Energy prospects

The numerical results examined in Section 4 have shown some


dependency trends of the rooftop velocity and skew angle as
a function of the main geometrical features of the installation
building and of its surroundings. Although the velocity variations
can be straightforwardly exploited to compare the different
geometrical solutions, some interesting considerations can also be
made upon examination of the energy perspectives connected to
a modified flow field over the building.

5.1. Energy potential

Seven configurations (Table 5) have been here compared in


terms of the Cubic Average Wind Velocity (u ) of the installation zone
(Eq. (4)), where fw is the frequency of the velocity class w in the
wind distribution). More specifically, the velocity variation in the
rooftop installation area with respect to the level in the urban wind
profile at the same height was calculated (Table 6) for the reference
wind velocities range from 3 m/s to 20 m/s (i.e. a suitable func-
Fig. 13. Mean velocity variation (a) and skew angle (b) in the analysis section at the
tioning range for a microeolic wind turbine); then, the net available
rooftop of the investigated buildings.
wind distributions have been calculated and the energy potential
per unit area (P) was evaluated (Eq. (5)) under two main assump-
buildings seems to represent a “second order” parameter to tions (see Refs. [7,8]):
consider with respect to the main structure proportions, its influ-
ence on the rooftop flow field must be carefully analyzed due to the - A prevailing wind direction was considered, in which all the
deviation effects induced on the streamlines. oncoming wind during the year was supposed to blow. This
Upon examination of Fig. 13, it is readily noticeable that the assumption is often verified in urban applications, in which the
geometric proportions between UB and IB have the most notable installation building is often chosen to ensure a façade
influence on the velocity variation at the rooftop of IB; Case 1 perpendicular to the prevailing wind;
configurations (with or without the sloped roof) show indeed - The wind in that direction was assumed to have a Rayleigh
a constant decrease of the flow velocity with respect to the velocity distribution, with the scale factor to match the
undisturbed wind profile, whereas Case 2 configurations constantly attended mean velocity of 6 m/s at Href.
offer a positive velocity variation, which is, however, maximized by
the application of the 8 sloped roof. This geometrical solution
seems in fact to provide the better flow field in the investigated vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u 20
u X
zone on the rooftop of IB also in Case 1, mainly due to the fact that u ¼ t
3
u3w $fw (4)
the flow is more guided over the roof and the recirculating zone in w¼3
the downwind zone is reduced (see also Ref. [8]).
Moreover, in Case 1 configurations a more pronounced depen-
dence from the inclination angle of the roof was found both in
Table 5
terms of velocity variation and of skew angle, i.e. the angle between Investigated configurations.
the local velocity vector and the wind main direction. This latter
IB height UB height Distance between UB width Inclination
parameter is of particular interest in the evaluation of an urban
(H) (h) IB and UB (D) (L) angle (c) ( )
installation site especially when Vertical Axis Wind Turbines
Case A 2.00 1.50 0.50 0.50 0
(VAWTs) are used, whose performance actually has a notable Case B 3.00 2.00 0.50 1.00 0
dependence on the oncoming skews angle (see Ref. [8]). Case C 4.00 2.75 0.50 1.00 0
In Case 2 configurations, conversely, a lower dependence from
the roof slope can be appreciated and almost the same skew angle Case C-I 2.75 0.50 1.00 8
was found with sloping roofs with either c ¼ 8 or c ¼ 18 . Case C-II 1.50 0.50 1.00 0
Case C-III 2.75 2.00 1.00 0
Finally, it is worth pointing out that, although the influence of Case C-IV 2.75 1.00 2.00 0
the sloped roof seems to be almost constant with respect to the
F. Balduzzi et al. / Renewable Energy 45 (2012) 163e174 171

Table 6
 ).
Calculated cubic average wind speed (u

 [m/s]
u Variation with respect to the
undisturbed distribution (%)
Case A 8.04 14.9
Case B 5.91 þ1.4
Case C 8.79 þ3.0
Case C-I 9.02 þ5.6 Fig. 14. Cases A, B and C geometries.
Case C-II 8.50 0.4
Case C-III 8.29 2.8 In particular, the skew angle can notably modify the power
Case C-IV 8.26 3.3
coefficient of the turbine: the choice of the turbine typology is
consequently significantly relevant in defining the final energy
harvesting. Within this context, an opposing behavior has been
1 highlighted by the technical literature for HAWTs and VAWTs.
P ¼ ru3 (5) According to the Wind Energy Handbook [35], Glauert gives the
2
thrust force of an actuator disk in yaw (Eq. (6)).
Fig. 14 reports a comparison between the geometrical configu-
rations A, B and C, which have been found to maximize the energy
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
cP ¼ 4a 1  að2cos g  aÞðcos g  aÞ (6)
content of the flow in the rooftop of IB for the three considered H
values: the P values of the investigated cases are reported in Fig. 15. This equation can be used to calculate the performance coefficient
Upon examination of Fig. 15, it is worth pointing out that: of a HAWT in skewed flow. According to Eq. (6), the optimum
performance coefficient of a HAWT shows a decrease for increasing
- The installation building height (H) has a capital relevance in angle g as shown in Fig. 18.
achieving a good wind energy potential in the rooftop zone (a Conversely, recent works on Darrieus VAWTs (Refs. [8,13,36])
building two times higher leads to a quadruple energy have shown that these rotors can indeed take some advantages
potential); from a functioning in skewed flow for g angles between 0 and 30
- In case of an IB height of 2, no velocity increment can be (i.e. the range of skew angles attended in the built environment).
obtained; Focusing on the results of Fig. 18, one can readily understand
- The impact of the building’s geometry in terms of available how the right choice of a microeolic turbine in the built environ-
energy is amplified for high H values; more specifically, when ment must be derived from the best compromise between the
a high installation building is considered, a suitable geometry energy potential in the installation site and the application of the
of the surrounding buildings can notably increase the pro- most suitable rotor for the specific conditions on the rooftop of the
ducibility of the wind turbine. investigated building.
In fact, where a HAWT ensures higher peak efficiencies, it
Focusing the attention on the H ¼ 4 solution, in Fig. 16 the conversely presents a more pronounced dependence on the skew
investigated C cases are compared and their P variations with angle: as a result, for high g angles, the energy harvesting can be
respect to the undisturbed condition are reported in Fig. 17. remarkably decreased. On the other hand, a Darrieus VAWT is
Fig. 17 leads to some interesting considerations. In particular, it deemed to improve its functioning in low and medium skew angles
is worth noticing that: and could provide a compensation to the lower peak efficiency.
In order to obtain a first estimation of the effect of a the skew
- The 8 skewed roof sensibly increases the energy potential in angle on the energy potential per unit area, a basic assumption was
the investigated zone on the top of IB with respect to the flat- introduced; in detail, the ratio between the power coefficient in
roof geometry (C-I vs. C); skewed flow and that in aligned flow conditions was assumed to be
- Despite a constant H value, an unfavorable h/H ratio can imply constant at each wind velocity. This assumption is quite well veri-
a notable reduction of the energy availability (C-II vs. C); fied for a wide range of oncoming wind, with the partial exception
- The global distance (L þ D) between the buildings has a great of the low-velocity conditions, where a stronger dependence of the
influence on the flow structure between the buildings them- aerodynamic coefficients of the blades from the Reynolds number
selves due to the discussed impact on the a angle; in particular, has to be considered.
if D exceeds a critical value, the “aerodynamic ramp” effect of
UB cannot be exploited anymore and the energy content of the
flow over IB is sensibly reduced (C-III and C-IV vs. C).
- With a constant Q ¼ L þ D, the energy content has however is
dependent on the L/D ratio, due to the different flow structure
which is established in the gap between the buildings (C-III vs.
C-IV).

5.2. Skew angle effects

The energy potential variations described in Section 5.1 are not


sufficient to define a realistic energy perspective of each solution.
Besides the unavoidable modifications induced by the real geom-
etry (i.e. obstacles, street furniture, variable roughness of the urban
context, etc.), one has also to take into account the effect of the
skew angle of the flow on the rooftop on the performance of the
machine, which is not included in the definition of the energy
potential. Fig. 15. P comparison (A, B and C cases).
172 F. Balduzzi et al. / Renewable Energy 45 (2012) 163e174

Fig. 16. Cases C geometries.

Fig. 19. P comparison (A, B and C cases) with skew effect.

Fig. 17. P variation with respect to the undisturbed condition e C cases.

On this basis, the impact of the skew angle as a function of the


turbine typology is highlighted by Figs. 19 and 20, where the
Fig. 20. P variation with respect to the undisturbed condition e C cases with skew
energy projections of Figs. 16 and 17 have been compared to
effect.
the new estimations for an HAWT or a VAWT installation,
respectively.
As one may notice, the skew effect has relevant impact on the 6. Conclusions
energy perspective of HAWTs installation, although the real energy
production could be partially compensated by an higher absolute In this study, a wide-ranging investigation on the flow modifi-
power coefficient. Conversely, VAWTs seem to provide very inter- cations induced on the rooftop area of a building in an urban
esting perspectives in terms of efficient exploitation of the flow context due to the surrounding environment has been presented.
which incomes on the rooftop of tall buildings, as they take great With this goal in mind, a parametrical CFD analysis was carried
advantage from the skewed-flow condition which actually out to characterize the flow field in a 2 m zone over the upwind
compensates their lower efficiency. corner of the selected building as a function of the installation
building height, the height and width of its upwind building and
the distance between the buildings themselves; in addition, the
presence of either a flat or a sloping roof was considered.
The parametric study pointed out some general tendencies,
which showed that the urban environment can indeed represent an
interesting frontier for a wide diffusion of the microeolic tech-
nology on the basis of the possibility to exploit some intense local
accelerations of the oncoming flow whenever particular geometric
conditions of the built scenario are fulfilled.
More specifically, in order to achieve some notable benefits in an
urban installation, the chosen building should be sufficiently higher
than the surrounding buildings. Moreover, the heights ratio between
the building and its upwind building in the prevailing wind direction
must be a function of their relative distance, in order to exploit a sort
of “aerodynamic ramp” effect which accelerates the flow oncoming
on the installation building; in particular, the angle between the
upwind corners of the two buildings is deemed to represent the most
influencing parameter on the flow behavior, which is, however,
Fig. 18. Variation of the power coefficient of the turbine induced by the skew angle. slightly modified also by the width of the upwind building. In
F. Balduzzi et al. / Renewable Energy 45 (2012) 163e174 173

addition, the roof shape can also provide some notable benefits in References
terms of flow direction and modulus, with a velocity increase
maximized by the application of an 8 sloping roof. [1] Mertens S. Wind energy in the built environment. Brentwood (UK): Multi-
Science Publishing; 2006.
Furthermore, with the assumption of a Rayleigh wind distri- [2] Dayan E. Wind energy in buildings: power generation from wind in the urban
bution, some specific configurations were also compared in terms environment e where it is needed most. Refocus 2006;7(2):33e8.
of specific energy potential, with the final results that a microeolic [3] Beller C. Urban wind energy e state of the art 2009. Report No.: Risø-R-
1668(EN). Roskilde (Denmark): Risø Laboratory e DTU; 2009.
turbine in the urban context can indeed theoretically take relevant [4] CSE Council. Ealing urban wind study. Final Report. Bristol (UK): CSE (Centre
advantages (up to 20%) from a favorable installation site, whereas for Sustainable Energy); 2003.
a very poor energy producibility has to be accounted for whenever [5] Syngellakis K. Urban wind turbines: development of the UK market. In:
Proceedings of the European wind energy conference. Athens (Greece);
an unfavorable environment is chosen. Finally, the prospects in 2006.
terms of of energy potential have been highlighted focusing on the [6] Banks D, Cochran B, Denoon R, Wood G. Harvesting wind power from tall
choice of both the installation site and the turbine typology. buildings. In: Proceedings of the CTBUH 8th world congress, Dubai (UAE),
2008 by Chicago. Council on Tall Buildings and Urban Habitat; 2008. p.
320e7.
Acknowledgments [7] Balduzzi F, Bianchini A, Carnevale EA, Chesi A, Ferrari L. Influence of the
building geometry on microeolic installations in the urban context. In:
Proceedings of world renewable energy congress XI. Abu Dhabi (UAE);
Thanks are due to Prof. Giovanni Ferrara of the University of 2010.
Florence for his support in this study. [8] Balduzzi F, Bianchini A, Carnevale EA, Ferrari L, Magnani S. Feasibility
analysis of a Darrieus VAWT installation in the rooftop of a building. In:
Proceedings of the international conference on applied energy (ICAE).
Glossary Perugia (Italy); 2011.
[9] Gryning SE, Batchvarova E. Modelling of the urban wind profile, air pollution
modeling and its application XIX. In: NATO science for peace and security
Acronyms series. Brussels (Belgium): Springer; 2008. pp. 18e27.
[10] Davenport AG. The response of six building shapes to turbulent wind. Philo-
ESDU Engineering Science Data Units sophical Transactions of the Royal Society of London, Series A: Mathematical
HAWTs Horizontal Axis Wind Turbines and Physical Sciences 1971;269(1199):385e94.
IB Installation Building [11] Cionco RM, Huber A, Tang W. Preliminary results of CFD simulations for the
scenario of a recent field study in an urbanized domain. In: Proceedings of
RSD Relative Standard Deviation the seventh symposium on the urban environment. San Diego (CA, USA);
SST Shear Stress Transport 2007.
VAWTs Vertical Axis Wind Turbines [12] Mertens S. Wind energy in urban areas: concentrator effects for wind turbines
close to buildings. Refocus 2002;3(2):22e4.
UB Upwind Building [13] Sharpe T, Proven G. Crossflex: concept and early development of a true
building integrated wind turbine. Energy and Buildings 2010;42:
2365e75.
Greek symbols
[14] Mertens S. The energy yield of roof mounted wind turbines. Wind Engi-
P energy potential per unit area neering 2003;27(6):507e17.
a angle between the corners of IB and UB [ ] [15] Ratti C, Di Sabatino S, Caton F, Britter R, Brown M. Analysis of 3-D urban
g skew angle [ ] databases with respect to pollution dispersion for a number of European
and American cities. Water, Air and Soil Pollution: Focus 2002;2(5e6):
3 turbulent kinetic energy dissipation rate [m2/s3] 459e69.
k Von Karman constant [16] Martin CL, Longley ID, Dorsey JR, Thomas JR, Gallagher MW, Nemitz E.
r air density [kg/m3] Ultrafine particle fluxes above four major European cities. Atmospheric
Environment 2009;43:4714e21.
c sloping angle of the roof [ ] [17] Engineering Science Data Unit. Strong winds in the atmospheric boundary
u specific turbulence dissipation rate [1/s] layer. Part 1. Mean-hourly wind speeds, ESDU 82026 with amendment A and
B. London (UK): ESDU; 1984.
[18] OpenCFD Ltd. Website, http://www.openfoam.com; 2011.
Latin symbols [19] Franke J, Hirsch C, Jensen AG, Krüs HW, Schatzmann M, Westbury PS,
Ah area occupied by the buildings [m2] et al. Recommendations on the use of CFD in wind engineering. In:
Proceedings of the international conference on urban wind engineering
Cm turbulence model constant and building aerodynamics. Sint-Genesius-Rode (Belgium): von Karman
CS roughness constant Institute; 2004.
D distance between UB and IB [m] [20] Franke J, Hellsten A, Schlünzen H, Carissimo B. Best practice guideline for the
CFD simulation of flows in the urban environment. Brussels (Belgium): COST
H
ˇ IB height [m]
Office; 2007.
H mean buildings height [m] [21] Tominaga T, Mochida A, Yoshie R, Kataoka H, Nozu T, Yoshikawa M, et al. AIJ
HR height of the roughness elements [m] guidelines for practical applications of CFD to pedestrian wind environment
around buildings. Journal of Wind Engineering and Industrial Aerodynamics
Ks sand-grain roughness [m]
2008;96(10e11):1749e61.
L upwind building width [m] [22] Blocken B, Stathopoulos T, Carmeliet J. CFD simulation of the atmospheric
Q global distance between the upwind corners of IB and UB boundary layer: wall function problems. Atmospheric Environment 2007;
[m] 41(2):238e52.
[23] Richards PJ, Hoxey RP. Appropriate boundary conditions for computational
Z maximum tested height of IB wind engineering models using the ke3 turbulence model. Journal of Wind
d displacement [m] Engineering and Industrial Aerodynamics 1993;46e47:145e53.
fu frequency of the velocity class u in the Rayleigh [24] Kim JJ, Baik JJ. A numerical study of the effects of ambient wind direction on
flow and dispersion in urban street canyons using the RNG ke3 turbulence
distribution model. Atmospheric Environment 2004;38(19):3039e48.
h UB height [m] [25] Heath MA, Walshe JD, Watson SJ. Estimating the potential yield of small
k turbulent kinetic energy [m2/s2] building-mounted wind turbines. Wind Energy 2007;10(3):271e87.
[26] Launder BE, Spalding DB. The numerical computation of turbulent flows.
u flow velocity [m/s] Computer Methods in Applied Mechanics and Engineering 1974;3(2):269e89.

u cubic average wind speed [m/s] [27] Cebeci T, Bradshaw P. Momentum transfer in boundary layers. New York
u* friction velocity [m/s] (USA): Hemisphere Publishing; 1977.
[28] Menter FR. Zonal two-equation keu turbulence model for aerodynamic flows;
w wind class
1993. AIAA paper 1993-2906.
yP height of the ground cells centroid [m] [29] Menter FR. Two-equation eddy-viscosity turbulence models for engineering
z0 roughness length [m] applications. AIAA Journal 1994;32(8):269e89.
174 F. Balduzzi et al. / Renewable Energy 45 (2012) 163e174

[30] Wilcox DC. Turbulence modeling for CFD. La Cañada (CA, USA): DCW Indus- [34] Leitl B, Shatzmann M. CEDVAL, compilation of experimental data for valida-
tries Inc.; 1994. tion of microscale dispersion model. Hamburg (Germany): Meteorological
[31] Eça L, Hoekstra M. Numerical aspects of including wall roughness effects in the SST Institute, Hamburg University; 1998.
keu eddy-viscosity turbulence model. Computers & Fluids 2011;40(1):299e314. [35] Burton T, Sharpe D, Jenkins N, Bossanyi E. Wind energy handbook. Oxford
[32] Ferrer E, Munduate X. CFD predictions of transition and distributed roughness (UK): John Wiley & Sons Ltd; 2001.
over a wind turbine airfoil. In: 47th AIAA aerospace sciences meeting. Orlando [36] Simão Ferreira CJ, van Bussel G, van Kuik G. An analytical method to predict
(FL, USA); 2009. the variation in performance of a H-Darrieus in skewed flow and its experi-
[33] Rakai A, Kristóf G. CFD simulation of flow over a mock urban setting. In: 5th mental validation. In: Proceedings of the European wind energy conference.
OpenFOAM workshop. Chalmers (Sweden); 2010. Athens (Greece); 2006.

You might also like