You are on page 1of 15

Proceedings of the ASME 2012 International Mechanical Engineering Congress & Exposition IMECE2012 November 9-15, 2012, Houston,

Texas, USA

IMECE2012-87139
DRAFT: A NUMERICAL INVESTIGATION OF IMPACT OF ARCHITECTURAL AND CLIMATIC PARAMETERS OF WINDCATCHER SYSTEMS ON INDUCED VENTILATION
Mohammad Mehdi Maneshi Department of Mechanical & Aerospace Engineering University at Buffalo Buffalo, NY-14260 Email: mmaneshi@buffalo.edu A. Scott Weber Department of Civil , Structural & Environmental Engineering University at Buffalo Buffalo, NY-14260 Email: sweber@buffalo.edu Amir Rezaei-Bazkiaei Department of Civil , Structural & Environmental Engineering University at Buffalo Buffalo, NY-14260 Email: ar92@buffalo.edu Gary F. Dargush Department of Mechanical & Aerospace Engineering University at Buffalo Buffalo, NY-14260 Email: gdargush@buffalo.edu
*

ABSTRACT The large energy demand from the HVAC industry for residential buildings, along with the ever growing need for the utilization of renewable sources of energy, has brought considerable attention to the area of induced ventilation. Windcatchers are green architectural structures historically used for the passive ventilation of the indoor spaces with the minimal non-renewable energy consumption. In this paper, a computational fluid dynamics (CFD) model is used to assess a windcatchers performance with different characteristics of the windcatcher design. For a single room with heat emitting objects, the effects of the windcatcher louver design and height were thoroughly evaluated in conjunction with a variety of dominant wind velocity and incident angles. A comparison was drawn between the ventilation efficiency for the cases of circular and rectangular windcatcher designs. Thermal discomfort (PD) values due to draft were calculated for different temperature and wind velocities. The developed model was employed to obtain the optimized configuration for the windcatcher-room system. The effect of ambient weather conditions on the results was investigated by performing the simulations for a range of air temperature and velocities. Results obtained in this paper provide windcatcher designers with valuable insights on the important design parameters such as windcatcher height, louver design and impact of ambient conditions.

INTRODUCTION Increases in the cost of conventional sources of energy and natural resource exploitation has pushed building designers and engineers to rethink the value of utilizing natural ventilation techniques. Naturally ventilated buildings tend to rely heavily on the ambient weather conditions and the flow patterns around and inside the building. Therefore, it is important for designers to: a) have a clear idea of how the air flow circulation and thermal performance of a naturally ventilated building are affected by the external parameters and b) understand the tools on the structural and architectural level to enhance or control the ventilation. Windcatcher systems are probably the oldest natural ventilation systems that have been used historically to cool the single and complex residential application in hot and arid regions [1], [2]. Due to the expensive experimental procedures for evaluation of natural ventilation mechanisms, computational fluid dynamics (CFD) modeling has gained extreme popularity in the past decade or two to further enhance their design. The enhanced computing power coupled with robust CFD modeling has become a reliable tool for building designers and architects to come up with more creative natural ventilation strategies for different applications [27]. This paper is focused on analyzing and evaluating the performance

of four different windcatcher louver designs for a small conditioned room via CFD modeling. Hughes et al., [2] investigated the effects of 8 different louver designs for rectangular windcatchers where the louver angle of 35 degrees demonstrated optimal performance. It was verified that the windcatcher is the most effective when placed right at the windward edge of the roof where the positive pressure is maximum and, the least effective at the leeward edge of the roof. Therefore, the ideal placement for multidirectional wind towers is at the center of the roof. It also was pointed that in spite of studies demonstrating better performance of rectangular cross-sections compared to other designs, most of commercial wind tower designs are circular, most probably due to the ease of manufacturing circular designs. In his Reynolds Averaged NavierStokes (RANS) analysis of effects of wind direction (incident angle) and velocity on cross-ventilation, Nikas found good agreement between the volume flow rate aerating and induced velocity profiles from the CFD model versus experimental results [8]. The study was centered around investigating the details of natural ventilation processes by closely looking at the shape and size of the vortices and the effects of building inner topology on the ventilation efficiency. A maximum inlet wind velocity of 5 m/s was used in the wind tunnel and CFD analysis which resulted in a maximum induced flow rate of 0.345 m3/s to the building for a zero-degree incident wind angle. The induced flow rates decreased with an increase in the incident wind angle. The numerical study confirmed that not only the building opening dimensions and incident angle but also the magnitude of wind velocity play an important role in the air exchange rates in the building. It also was concluded that, although the inner geometry of the building does not affect the average air-change rate, it has considerable effect on refreshing the air in inner regions of the building envelope. In our study, the selection of inner columns dimensions and attention to the form and size of formed vortices around the windcatcher system were inspired by Nikass study. The effects of the wind incident angle and magnitude on the quality of air exchange rate obtained form a windcatchers performance also have been confirmed in other similar studies [9], [10]. Li modeled a 5050 cm square windcatcher with incident wind velocities in the range of 0.6-5 m/s and four different incident angles. It was concluded that the induced flow velocity to the building increases with increase in the inlet velocity. Slight decrease in the induced flow due to increase in incident angle was reported for wind speeds below 3 m/s. The maximum induced flow rate was reported for the wind speed of 6 m/s in zero-degree incident wind angle to be 0.19 m3/s [9]. Elmualim employed an experimental wind tunnel and CFD modeling to compare the performance of four-section circular windcatcher systems to rectangular ones [11]. Pressure coefficient distributions, internal air speed and volumetric air flow rate in the test room were measured for the sake of comparison with the CFD model. Similar to the previous

studies, increases in induced flow rates was observed with an increase in wind velocity whereas a decline in induced flow rates occurred with increasing incident angle from zero to 45 degrees. A maximum normalized induced flow rate of 2.5 m3 /m2.s was reported through the rectangular design for the inlet velocity of 6 m/s in zero-degree wind attack angle versus approximately 0.5 m3/m2.s for the circular design. The flow catching efficiency of the rectangular design over circular was more pronounced for higher inlet wind velocities. A ventilation rate of 5 air changes per hour with the air velocity of 3 m/s was achieved with the circular windcatcher. Experimental wind tunnel and smoke visualization methods were employed by Elmualim [12] to investigate the effects of air flow control mechanisms and heat source inside the room on the windcatcher performance. A 5050 cm rectangular cross-section windcatcher of 150 cm height was analyzed. It was suggested based on the findings that the level by which the windcather system lowers the indoor temperatures in the presence of the heat sources is mainly dependent on the outdoor air temperature values. Maximum internal temperature observed in the presence of heat sources was 30-32 C when the external temperature was in the range of 20-22 C. The introduction of the heat source inside the room increased the induced flow rates specially for the cases with lower outside air wind velocities. With an 8 C temperature difference between internal and external temperatures, the windcatcher was capable of reducing the indoor temperature by approximately 6 C. It was concluded that the windcather installation was appropriate specially for overheated summer days. In another study, the efficiency of installed windcatchers was compared to the openable windows in identical conditions which resulted in considerably higher indoor induced flow rates with windcatchers on the roof versus windows on the wall [13]. In their analysis of a one-sided wind tower, Montazeri and Azizian studied the effects of placing objects, shorter and taller than the tower opening size, upstream [14]. Based on the wind tunnel analysis, they found that placing a short upstream object before the windcatcher reduces the circulation area thus increases the effective inlet area that consequently contributes to higher performance levels. A taller upstream object on the other hand, leaves the tower in the wake region of upwind where the tower works like a suction device. In another study, an experimental and CFD analysis were performed on a twosided windcatcher system in a channel flow experiment. The stack effects of a one-sided systems was compared to a twosided system which resulted in less fluctuant performance levels achievable with two-sided system in different wind speeds and incident angles [15]. Montazeri [16] carried out another study on the effects of openings and incident wind angles on windcatchers performance with an array of two to twelve-sided windcatchers in a wind tunnel. It was concluded that the induced air flow rate decreases with increase in the number of openings. However, it was suggested based on the results that the sensitivity of the windcatchers performance to the incident wind angle decreases with increasing number of openings. Montazeri also stated that

the efficiency of the windcatchers performance strictly depends on creating maximum pressure difference between the inlet opening and exhaust. It also was confirmed that it is the air movement around the building that determines the optimal placement of the tower and opening sizes for the best performance. Moreover, Montazeri concluded that the rectangular windcatcher exhibites a better efficiency than the circular one for the range of wind speeds and incident angles studied. Kalantar took into account the effects of the wind tower height, wall material properties, wind velocity, humidity, ambient and inside air temperatures to assess the cooling efficiency of an evaporative cooling tower [17]. It was demonstrated that the modeled cooling system is capable of reducing the indoor temperature by as high as 10-15 C and the main area of influence exists in the first 2 meters of the towers height. Liu [18] considered a number of layers and lengths of louver designs in a CFD model to study the effectiveness of the windcatcher design in a typical office building for efficient contaminant removal. It was confirmed by the results that the maximum airflow is achieved as a result of enhanced resistance and the short circuit at top louvers. The windcatcher exhibited

the highest ventilation rate when the louver length was equal to the reference length. Liu concluded that using a leeward side opening as a complement to the windcatcher design enhances the concentration removal and flow pattern uniformity. STUDY APPROACH AND ANALYSIS Except for a small number of studies focused on the indoor air quality and thermal comfort achieved by the adoption of windcatcher systems in residential buildings, the majority of the research has focused on building a knowledge inventory of flow patterns and accuracy of the CFD model predictions compared to limited experimental results. From this previous work, variation in ambient air temperature, humidity levels, building floor dimensions and height, interaction between different stories, inducing effects of neighboring buildings, and the amount of solar radiation absorbed by the building can potentially alter the obtained results in the literature to a considerable extent. The central focus of this study has been put on evaluating the performance of two different louver designs for both rectangular and circular windcathers in a modeled room via CFD analysis by controlling three main parameters; wind speed, air temperature and architectural features of the windcatcher.

FIGURE 1. Windcatcher-room system schematic with heat emitting objects

Physical Domain Studied In this research, the ventilation efficiency of a windcatcher has been studied for a simple room model. The room modeled is a single story building with the only opening being the windcatcher system in the middle of the roof. To consider the effects of indoor human occupancy and equipment, it was assumed that there are two cylindrical (individuals) and cubic (computers) heat sources in two sides of the room (Figure 1). These

objects are assumed to generate a constant heat flux of 450 W/m2. The only architectural elements considered in the room are two columns of 20 by 20 cm, on the side walls with 10 cm protrusion from the wall surface which are similar to the values used by Nikas [8]. A schematic of the room-windcatcher system is depicted in Figure 1. Windcatcher Design The most well-studied physical aspects of windcatcher design are the size and number of

openings [1416] and louver design [18]. Louver design has been focused mainly on the impact assessment of number of openings on It is noteworthy that the combination of these physical parameters and ambient weather conditions can severely affect the windcatcher performance predictions, a matter that has not been thoroughly addressed in the literature at this time.

Louver designs with 5 and 10 layers were selected for comparison for both rectangular and circular windcatchers (cases I and II versus III and IV). To analyze the performance of each of these louver designs, a range of incident wind speeds and attack angle were analyzed in the initial tier of design to obtain the near-optimal physical configuration for the windcatcher. Incident wind speeds of 5, 10, 15, and 20 m/s were used to cover a vast range of common windcatcher design conditions. The net induced air flow rate entering the building was considered as the main criteria for the comparison of different windcatchers and their efficiency in catching the outer flow. Incident wind angles of 0 and 45 degrees were selected for comparison. First, the model was run for taller rectangular windcatcher designs (cases I and II) with the combination of different wind speeds, incident angle and louver designs. The initial results for rectangular system were then compared to the runs for the circular windcatchers of lower height (cases III and IV) to compare the differences with the change in the windcatcher height. Followed by the comparison between the ventilation efficiency of the first four cases, with successively decreasing height, a series of simulations were performed for the case V, the circular design of the same height as case I, to provide a comparison between the performance of different cross-sections with the same height. The results from this stage were analyzed for the inlet mass flow rate capturing capacity to select the choice design to analyze the indoor air quality performance by considering the effects of radiation model and occupant internal energy production rates. CFD MODEL To construct a well-defined CFD model and for ease of grid generation, the whole fluid field was divided into three domains. The first domain consisted of the room, as defined in Figure 1, in which further thermal model analysis were conducted. The second and third included a small domain surrounding the windcatcher, and the ambient air surrounding the entire system. These domains were defined on the basis of the room hydrodynamic length scale, H, which is 8 meters. The ambient air inlet plane is 3H from the leading edge of the room and the outlet is 5H downstream from the rooms trailing edge to ensure uniform flow further downstream. The far-fields, i.e. the air surrounding the top of the roof and the room side walls, were considered to be 4H apart to be fairly undisturbed from the viscous effects induced by the room in the flow field. Because the problem is symmetric, only one half of the solution domain was modeled and solved. To numerically solve for the flow field, a very smooth tetragonal grid was generated with the goal of

FIGURE 2. Geometrical parameters of windcatcher design

The focus of this study has been put on evaluation of a combination of physical and meteorological parameters. We chose to analyze three physical parameters; cross-section (rectangular versus circular), windcatcher height (1.9 to 1.18 m) and louver design (5 versus 10-level) and two main ambient weather characteristics; wind speed and ambient air temperature through the course of this study for a single room. A list of different design scenarios (cases) in terms of geometrical configuration of the windcatcher is presented in Table 1. Theta is the louver inclination angle equal to 45 degrees for all designs, L3 is the width/diameter of the windcatcher cross section equal to 50 cm, and t is the opening gap for each louver in the windcatcher which is equal to 4.3 cm. For a description of each of the dimensions refer for Figure 2. Each parameters significance will be further studied in the results section.
TABLE 1. Dimensions of different windcatcher designs
Case Louver level Cross section L1 (cm) L2 (cm)

I II III IV V

5 10 5 10 10

Rect. Rect. Circ. Circ. Circ.

118 140 68 90 140

168 190 118 140 190

placing the nodes more densely near solid surfaces, especially the windcatcher and near high ventilation zones such as windcatcher louvers and interfaces to the room and etc. To model near wall behaviours of the fluid and to meet the criterion, , a boundary layer grid was generated to ensure a minimum of 5 nodes inside the viscous boundary layer. The final mesh generation around the windcatcher is depicted in Figure 3.

[19] are appropriately credited with developing the standard model, with Launder and Sharma [20] providing improved values of the model constants. From these two quantities can be formed a length scale ( ), a timescale ( ), a

quantity of dimension ( equations for k and epsilon are:

, etc. The transport

(2)

(3)

FIGURE 3. Grid formation around the windcatcher

Where: The induced flow pattern and temperature distribution inside the room and outside are governed by the general transport laws and conservation of mass, momentum and energy. The model applied here, includes the numerical capability for solving the continuity, Navier-Stokes, and energy for turbulent flows. The general form of all possible transport properties obey the Eqn. (1) in which is the considered variable and are the velocity components in direction and is the fluid density.

, In which is the coefficient of thermal expansion, and is the turbulent Prandtl number for energy. For the standard and realizable models, the default value of is 0.85, and the constants are empirical values calibrated for most compatibility of turbulence regimes.

(1)
The Boussinesq Buoyancy Approximation The basis of this approximation is that there are flows in which the temperature varies little, and therefore the density varies little, yet in which the buoyancy drives the motion. Thus, the variation in density is neglected everywhere except in the buoyancy term. If denotes the density at the reference position where temperature is , for small temperature gradient inside the fluid we can write

Where is the source term and term related to a variable . The Model

is a constant

The model belongs to the class of twoequation models, in which model transport equations are solved for two turbulence quantities - i.e., and in the model. The model is the most widely used complete turbulence model and it is incorporated in most commercial CFD codes. As is the case with all turbulence models, both the concepts and the details evolved over time; but Jones and Launder

The buoyancy term is the same order of magnitude as the inertia, and the acceleration or the

viscous stress and is not negligible. Being so, the buoyancy term can be written as

Boundary Conditions and Assumptions The boundary conditions applied to the evaluated windcatcher are discussed here. The far-field flow boundary condition was set to free-slip wall in CFX reflecting the assumption that the flow field in infinity is not disturbed by the viscous effects of the solid boundaries of the room and windcatcher. The outlet boundary condition, being far enough from the room, i.e five times the room dimensions, was set to the static pressure of zero. All wall boundary conditions were set to no-slip, adiabatic and gray diffuse. In the first tier of simulations, the inlet air temperature was assumed to be 22C while later on it was changed to study its effect on the windcatcher ventilation efficiency and occupant thermal comfort. Radiation Model To fairly model the local solar heat transfer and indoor and outdoor radiation effects on the heat transfer, the position of the sun in the sky must be known. Besides, the intensity of incident total solar radiation and the thermal radiation exchange between each pair of all the possible surfaces must be taken into consideration. This results in a source term that is added to the energy equation. In the present work, we used an average solar radiation intensity of 1360 W/m2 . Total solar radiation incident on a sloped surface inside the domain is: (4) Where is the total incident radiation, is the direct solar radiation, the sky diffuse radiation, and is the ground reflected radiation to the identified surface. Shortwave radiation incident upon a surface was considered to be converted to a thermal source, depending on the local absorption coefficient of the material. The fluid mechanic solver of ANSYS CFX offers four radiation models; Discrete Transfer Radiation Model; P-1 Radiation Model; Rosseland Radiation Model; and Monte Carlo Radiation Model. From our previous studies it was noted that Discrete Transfer Radiation Model is more suitable for natural ventilation simulations. The main assumption applied in the DTRM model is that radiation leaving surface element in a specific range of solid angles can be approximated by a single ray. It uses a ray-tracing algorithm to integrate radiant intensity along each ray and is a relatively simple model that increases accuracy by increasing the number of rays over a wide range of optical thicknesses.

In this study, as mentioned before, we used discrete transfer method while considering air as a non-participating media and using eight rays to confirm high accuracy. All surfaces are considered to be gray and diffusive. The emissivity for human skin was assumed to be 0.95, for oil paints of the room walls, 0.9 and for the computer 0.98 [21]. Grid Study To make sure that the results obtained in the CFD analysis were not grid dependent, several grid sizes were tested. For the case of 10-level rectangular and circular windcatcher a total of 1478000, 2575000, 3730000 and 5160000 elements were tested where obtained results were approximately 1.86% different in the worst case. A total mesh size of 3730000 elements was utilized for the rest of simulations to ensure fine convergence and high numerical accuracy. The numerical analysis was performed in a twotiered fashion. The first phase mainly focused on the windcatcher flow characteristics and inlet air capturing capacity, which are the main criteria for performance evaluation. The first phase provided the best design that was subsequently utilized to model the mass flow rates harvested from the inlet flow, and indoor temperatures to evaluate the achievable comfort levels in a range of ambient conditions in phase 2. The details of the physical problem definition, windcatcher design, and numerical methods used are described in this section. RESULTS AND DISCUSSION Rectangular versus Circular Design Preliminary simulations of the CFD model were designed to assess and compare the performance of circular and rectangular designs with respect to the range of inlet wind speeds and catcher height in a range of incident wind speeds and angles. The cross-section of each of the rectangular and circular designs has four compartments or sections (A, B, C, and D). As depicted in Figure 4 to Figure 7, these sections act as inlet/exhaust to/from the building depending on the wind incident angle and crosssection design. After each run of the roomwindcatcher system model, the magnitude of the mass flow rate at the bottom of each of the four sections of windcatcher was calculated to provide a criterion for comparing the efficiency of rectangular and circular systems. A summary of the induced mass flow rates (kg/s) for the first four design cases is presented in Table 2 and Table 3. Positive values in Table 2 and Table 3 indicate that the flow is directed toward the room whereas negative values indicate the induced flow exiting the

room through the specified windcatcher cross-section. As it is depicted in Figure 4 and Figure 6, crosssections A and B are windward for the incident angle () of zero whereas cross-sections B and D would be leeward. In the case of incident angle of 45 degrees, flow stream hits the windcatcher first from crosssection A where cross-sections B and D are on the lateral sides and section D is on the leeward side (Figure 5 and Figure 7). A close examination of Table 2 for rectangular system shows that air enters the room from windward sections (A and B) and exits the room from the leeward sections (C and D) in the case of zero-degree incident angle. This effect is demonstrated in a sample velocity snapshot depicted in Figure 4. This pattern repeats itself for all wind strengths with expectedly higher induced flow rates to the room for higher incident wind speed. The comparison of mass flow rates captured by rectangular 10-level louvers

compared to 5-level louvers shows that the 10-level design is capable of capturing almost twice as much flow as 5-level system for all wind speeds with zero degree incident angel. For 45-degree incident angle, there were changes in the system behavior as the wind speed increased. The majority of the inflow to the room for the wind speed of 20 m/s in the 5-level louver rectangular systems enters from the leeward cross-section (C) whereas this pattern for lower wind speeds is reversed at 15 m/s or has a more balanced distribution between the windward and leeward sections at 10 and 5 m/s. This behavior occurs primarily because of the difference in the size of vortices formed at the louver level on top of the 10-level design versus 5-level as depicted in Figure 8 and Figure 9, respectively.

TABLE 2. Mass inflow rates (kg/s) to the room at the bottom of each section of rectangular windcatcher in different wind speeds Rectangular windcatcher V (m/s) 20 20 15 15 10 10 5 5 Incident angle = 0 = 45 = 0 = 45 = 0 = 45 = 0 =45 5-level louver sections (case I) A 0.1544 0.0276 0.1197 0.1236 0.0818 0.0822 0.0458 0.0226 B 0.1544 -0.0732 0.1197 -0.0657 0.0818 -0.0452 0.0458 -0.0215 C -0.1544 0.1178 -0.1197 0.0072 -0.0818 0.0082 -0.0458 0.0202 D -0.1544 -0.0732 -0.1197 -0.0657 -0.0818 -0.0452 -0.0458 -0.0215 Net inlet mass flow 9100.8 9100.8 6825.6 6825.6 4550.4 4550.4 2275.2 2275.2 10-level louver sections (case II) A 0.3929 0.5934 0.2941 0.4216 0.2046 0.3124 0.1098 0.174 B 0.3929 -0.2068 0.2941 -0.1526 0.2046 -0.107 0.1098 -0.0572 C -0.3929 -0.1798 -0.2941 -0.1164 -0.2046 -0.0984 -0.1098 -0.0596 D -0.3929 -0.2068 -0.2941 -0.1526 -0.2046 -0.107 -0.1098 -0.0572

TABLE 3. Mass inflow rates (kg/s) to the room at the bottom of each section of circular windcatcher in different wind speeds Circular windcatcher V (m/s) 20 20 15 15 10 10 5 5 Incident angle = 0 = 45 = 0 = 45 = 0 = 45 = 0 =45 5-level louver sections (case III) A -0.1709 -0.1062 -0.1112 -0.0679 -0.0696 -0.056 -0.0354 -0.028 B -0.1709 -0.0195 -0.1112 -0.0181 -0.0696 -0.0099 -0.0354 -0.0063 C 0.171 0.145 0.1112 0.1042 0.0696 0.076 0.0354 0.0406 D 0.171 -0.0195 0.1112 -0.0181 0.0696 -0.0099 0.0354 -0.0063 Net inlet mass flow 9100.8 9100.8 6825.6 6825.6 4550.4 4550.4 2275.2 2275.2 10-level louver sections (case IV) A -0.1954 -0.0612 -0.1321 -0.0438 -0.08 -0.037 -0.0415 -0.0202 B -0.1954 -0.0503 -0.1321 -0.038 -0.08 -0.0231 -0.0415 -0.0126 C 0.1954 0.1618 0.1321 0.1198 0.08 0.0834 0.0415 0.0456 D 0.1954 -0.0503 0.1321 -0.038 0.08 -0.0231 0.0415 -0.0126

FIGURE 4. Velocity field for zero degree incidence rectangular windcatcher in 10 m/s incident wind velocity

case where the flow enters from the windward sections (A and B). The location and size of the wake regions formed in front and behind the circular design (look at sample flow patterns depicted in Figure 11 and Figure 12) causes a partial windward and leeward flow capture by the circular design. The location of the windcatcher in the middle of the roof gives the system design a more balanced tone to increase the capacity to benefit from vortices formed on both sides of the roof. The increasing trend of induced inflow rates to the room with increasing wind speed is similar to the rectangular system for both 5 and 10-level louver designs.

FIGURE 6. Velocity field for zero degree incidence circular windcatcher in 10 m/s incident wind velocity

FIGURE 5. Velocity field for 45-degree incidence rectangular windcatcher in 10 m/s incident wind velocity

The width of the vortex formed behind the 5-level louver becomes larger and more visible than the 10level design so that the flow enters the room from the leeward section as louver falls right at the tale of the formed wake Figure 9. The similar comparison for the 10-level louver design reveals that the larger height of the windcatcher middles with the wake region formation behind the windcatcher so there is no entrance from leeward section and all induced flow rate is caused by the windward vortex formed by the front side edge of the room (Figure 8). The magnitude of mass flow rates to the building for the 10-level louver is expectedly higher than the 5-level louver design. Although the 45-degree incident angle mass flow rate values are comparatively lower than the zero-degree, the values for the 10-level louver design are much higher than the corresponding values for the 5-level design for the 45-degree incident wind angle. A number of observations were obtained by comparing the corresponding values in Table 3 for the circular designs to the values for rectangular one for comparatively lower heights. The entering flow sections for all zero degree incident angles are the leeward (C and D) ones as compared to the rectangular

FIGURE 7. Velocity field for 45-degree incidence circular windcatcher in 10 m/s incident wind velocity

It can be concluded from the simulation results that there is an optimal placement that guarantees maximum wind capture. This position is dependent on the position of the windcatcher on the roof, its height and opening size (louver design in this case). There is a minimum height before which not much of the inflow is captured because of the vortices formed in front of the building. As the height of the windcatcher increases, its effect on the shape and size of the formed vortices becomes more dominant so that higher portion of the inlet flow can be captured by the system. This observation is in agreement with previous studies [16].

In the case of 45-degree incident angles, the shorter circular design exhibits a different behavior than the rectangular one. The majority of induced flow rate to the room enters from the leeward section (C) only whereas the case for rectangular system that both leeward and windward sections (A and C) are involved (compare Figure 5 to Figure 7). A closer look at the flow patterns and formed vortices for the 45degree incident angles in Figure 8 to Figure 11 shows that the main contributing factor in determining how the induced air enters the rooms is a combination of windcatcher height and vortex height around the windcatcher. A similar behavior in the case of 10-level circular cross-section was observed where the leeward

section (C) is the main contributor to the inflow to the room. To examine the behavior of different designs, a comparative parameter was defined; normalized catching efficiency (NCE). NCE is equal to the percent mass flow rate captured by the windcatcher system compared to the inlet mass flow rate. Thus, NCE is equal to the percent ratio of sum of the induced flow rates to the building in each case (kg/s) to the net inlet mass flow rate (calculated from the inlet wind velocity and solution domain dimensions, kg/s) from Table 2 and Table 3. The NCE values were calculated for rectangular and circular designs then plotted versus inlet wind speed values.

FIGURE 8. Flow patterns for 10-level rectangular windcatcher in 10 m/s wind speed and 45-degree incident angle (at symmetry axis)

FIGURE 9. Flow patterns for 5-level rectangular windcatcher in 10 m/s wind speed and 45-degree incident angle (at symmetry axis)

FIGURE 10. Flow patterns for 10-level circular windcatcher in 10 m/s wind speed and 45-degree incident angle (at symmetry axis)

FIGURE 11. Flow patterns for 5-level circular windcatcher in 10 m/s wind speed and 45-degree incident angle (at symmetry axis)

FIGURE 12. Flow patterns for 5-level circular windcatcher in 10 m/s wind speed and zero degree incident angle (at symmetry axis)

10

FIGURE 13. Flow patterns for 10-level circular windcatcher in 10 m/s wind speed and zero degree incident angle (at symmetry axis)

The NCE values obtained for the 5 and 10-level louver rectangular designs (cases I and II) are plotted in Figure 14 for both zero and 45-degree incident angle versus modeled wind speed range. NCE was assumed to be a representative criterion for how efficient a windcatcher system is to capture the available air flow around the building. Results for the rectangular system in Figure 14 show that the 10-level louver design is able to capture the highest ratios of available air flow in both zero and 45-degree incident angles. The rectangular system exhibits better performance in zero-degree incident angle than 45-degree. This effect is more pronounced for the 10-level system. One interesting observation from this graph is that the air capturing capacity of the rectangular system decreased with increasing wind speed with a more highlighted effect for 10-level louvers. This can be interpreted as comparatively higher susceptibility of the 10-level louver design to wind speed variation. Nevertheless, the NCE of the 10-level louver is close to twice that of the 5-louver design for most wind speeds. The NCE values obtained for the 5 and 10-level louver circular designs (cases III and IV) are plotted in Figure 15 for both zero and 45-degree incident angle versus modeled wind speed range. Similar to the rectangular design, the 10-level louver in zero-degree incident angle demonstrates the highest performance compared to other circular designs in terms of capturing the inlet flow. There are differences in the patterns compared to rectangular design. The second best NCE is achieved by the 5-level louver design in zero degree incident angle rather than the 10 louver design in 45-degree, which is the case in rectangular windcatcher. This behavior of the circular design can be explained comparatively by looking at the flow patterns in front and behind the 5-level circular design (case IV) in zero-degree incident angle (Figure 12) and 10-level design (case III) in 45-degree incident angle (Figure 10). Even though the size of vortices formed in front and behind the 10level design seem slightly larger than the 5-level design, it seems from the flow patterns that the flow shoots off the entrance further for the 45-degree attack angle causing weaker wake region formation behind the building. As a result, the

induced flow rate from the leeward section (C) decreases in the case of 10-level circular design in 45-degree wind incident angle (Table 3) for the range of wind velocities.

FIGURE 14. Rectangular windcatcher NCE values versus wind speed in different incident angles (Cases I and II)

FIGURE 15. Circular windcatcher NCE values versus wind speed in different incident angles (Cases III and IV)

11

These results can be translated into higher sensitivity of the circular windcatcher to the incident wind angle than rectangular designs. Similar conclusions were drawn previously for circular design considering different number of openings in relatively smaller incident wind speeds [16]. It is thought that the sharp edges of the rectangular design create larger flow separation regions and higher pressure differentials across the windcatcher system [11] that makes the rectangular designs less susceptible to incident angle. Modifications to the circular windcatcher exterior, where the system is exposed to the air, can result in more flexible circular designs.

wind speed of 10 m/s for a range of inlet ambient air temperature values of 17, 22, 27, and 32 C to obtain velocity and temperature distribution in the room. The simulation was performed subsequently for a constant inlet air temperature of 17 C and a range of inlet air velocities; 20, 15, 10, and 5 m/s to cover a range of climatic conditions. The risk of discomfort due to draft was calculated based on Percent Dissatisfaction (PD) factor (Eqn. 5) to compare the thermal comfort condition for a range of ambient weather conditions [22]: (5)

where the main three parameters for evaluation of PD: velocity (V), m/s, air temperature (Ta), C, and turbulence intensity (Tu), were obtained from the CFD results on an average basis for a range of cross-sections across the rooms elevation. The elevation zero is the room floor and the points continue to the ceiling level. Average values of these parameters were obtained for calculation of PD in the room profile. To provide a sample of input parameters that led to calculation of PD, a snapshot of the temperature and velocity distribution at the occupants respiratory level (1.8 m) for wind speed of 15 m/s are plotted in Figure 17 and Figure 18, respectively. As it is evident from Figure 17 and Figure 18, the velocity values right below the windcatcher can be as high as approximately 6-7 times the average values in the rest of the room whereas the temperature distribution across the room is quite uniform except around the internal energy sources. This observation can be interpreted as higher dependency of local dissatisfaction levels on induced flow velocities than temperature values.

FIGURE 16. Comparison between NCE values of 10-level louver rectangular and circular designs of different height

After collecting information on the performance of different designs based on cross-section and height, one last set of simulations was performed with the configuration described as case V in Table 1. These runs were designed to compare the performance of windcatchers of the same height and louver design but different cross-sections (case II and V) to ensure selection of the most efficient design. The corresponding NCE values for the best circular and rectangular designs, cases II, IV and V, are presented in Figure 16. Each case has a 10-level louver design where the cases II and V are the tallest ones at 190 cm and case IV is at 140 cm height. All the NCE values are plotted for the zero-incident angle case to compare the effects of height and cross-section. The rectangular design (case II) shows higher NCE compared to the circular design (case V) of the same height. Making the comparison between circular designs, the taller circular design (case V) shows higher potential to capture the inlet flow rate than the shorter circular design (case IV) with the same louver design. Effects of Ambient Temperature and Wind Speed on Indoor Ventilation The results discussed in the earlier section led to the selection of the choice design; a rectangular 10-level louver design in zero-degree incident angle (case I). This configuration was employed to run the model for a range of temperatures and wind speeds for calculation of indoor occupant thermal comfort levels. The simulation was performed first for a single inlet

FIGURE 17. Selected temperature snapshot at respiratory level (averaged values used for calculation of PD valued)

The PD for the range of inlet temperature values in selected elevations in the room are depicted in Figure 19. Similar data for the range of simulated wind velocities were obtained for the constant inlet temperature of 17 C as depicted in Figure 20.

12

year to ensure an acceptable frequency of occurrence for desirable occupant comfort.

FIGURE 18. Selected velocity snapshot at respiratory level (averaged values used for calculation of PD valued)

The overall Figure 19 and Figure 20 shows an increase in PD values closer to the floor and ceiling whereas the corresponding values for the individuals activity levels (0.6, 1.2 and 1.8 m) are considerably lower. Higher induced flow velocity at floor and ceiling levels due to windcatchers functionality is responsible for high dissatisfaction at extreme high and low elevations in the room. It is noteworthy that the temperature difference between the floor and ceiling in all the cases was not more than 1 C to cause more discomfort due to temperature difference at head and toe level. PD values in Figure 19 are calculated based on a constant 5 m/s wind velocity for all cases to represent a low turbulence condition. It is observable from Figure 19 that the lower the inlet flow temperature values are the higher the percentage dissatisfaction values by the occupants will be. Nevertheless, the dissatisfaction due to draft at occupants activity levels are considerably lower (below 4%) for the entire range of inlet temperatures. The trend for higher velocity levels in Figure 20 are similar but with considerably higher magnitudes than the case with increasing inlet temperature. The PD values are consistently higher at floor and ceiling level as expected with PD values being in the range of approximately 15 to 30% as compared to the case with varying inlet temperature values. With higher wind velocities, higher dissatisfaction levels are shown to occur. This result was mainly due to draft and it increased almost linearly as a result of the direct relationship between the wind magnitude and draft dissatisfaction. The PD values for the wind velocity of 20 m/s show that about 20% of occupants are expected to express dissatisfaction whereas the case with 5 m/s wind that the PD values are about only 5% at occupants activity level. These results can be translated into relatively high dissatisfaction levels in comparatively average ambient temperature (17 C) values and high air velocities (20 m/s). Designers of natural ventilation systems should be mindful of these considerations while analyzing the dominant meteorological data of a prospective region for the worst case annual wind and temperature variations. These dissatisfaction level observations should be checked throughout the design

FIGURE 19. PD values versus elevation for selected simulations with varying inlet temperature values

FIGURE 20. PD values versus elevation for selected simulations with varying inlet wind velocities

SUMMARY AND CONCLUSIONS A windcatcher-room system was simulated via a CFD model to evaluate the effects of three main parameters; wind speed and incident angle, inlet air temperature, and architectural features of the windcatcher design. The modeling was performed in two steps. First, the effects of different louver design and windcatcher heights were simulated for a variety of wind speeds in zero and 45-degree incident angles. The ability of the windcatcher design to capture the inlet air flow rate was chosen to compare different design scenarios. An optimal design was selected for the second round of simulations to take into account the effects of internal sources of energy production for a variety of inlet air temperature and wind speeds. The importance of windcatcher height on the efficiency of the design was verified. It should be noted that most research undertaken in the natural ventilation literature focus on simple building geometries due to limitations such as the need for comparison to experimental data of limited availability. The possible missing link in such analyses therefore would be their inability to consider the effects of architectural elements of the real-world prospective building and the surrounding buildings.

13

This might be a fertile area for future work on natural ventilation to evaluate the ventilation efficiency in multi-zone buildings and to calibrate the urban scale effects of structures on neighboring buildings in terms of natural ventilation. It is suggested that a close examination of the effects of the building dimensions (floor dimensions) to windcatcher height should be performed for any particular design. This requires a clear understanding of the air flow regime around and above the building and windcatcher system in different incident angles to ensure the optimal configuration of the windcatcher for dominant climatic condition of a design site.

A summary of the conclusions from this study are: A rectangular cross-section system demonstrated higher efficiency than the circular one in different wind speeds, incident angle and heights. Windcatcher system generally demonstrated better performance in zero-degree wind attack angle. The rectangular design with 10-level louver demonstrated the best overall performance in zero-degree wind incident angle. Percent dissatisfaction due to draft in the room was considerably lower for the rectangular 10-level design in low wind velocities and inlet temperatures. Circular windcatchers had a relatively higher sensitivity to the incident wind angle than rectangular designs. 10-level louver rectangular design demonstrated slightly higher susceptibility to wind speed variation. There is a strong relationship between the windcatcher ventilation performance and the building floor dimensions. The optimal height is a more substantial parameter in windcatcher design than the number of openings with a similar louver design. The thermal dissatisfaction due to draft in the building can be as high as about 30% for higher wind velocities and lower inlet temperature values. References: [1] O. Saadatian, L. C. Haw, K. Sopian, and M. Y. Sulaiman, Review of windcatcher technologies, Renewable and Sustainable Energy Reviews, vol. 16, no. 3, pp. 1477-1495, Apr. 2012. [2] B. R. Hughes, J. K. Calautit, and S. A. Ghani, The development of commercial wind towers for natural ventilation: A review, Applied Energy, vol. 92, pp. 606-627, Apr. 2012. [3] Y. Jiang, D. Alexander, H. Jenkins, R. Arthur, and Q. Chen, Natural ventilation in buildings: measurement in a wind tunnel and numerical simulation with large-eddy simulation, Journal of Wind Engineering and Industrial Aerodynamics, vol. 91, no. 3, pp. 331-353, Feb. 2003. [4] S. Hussain and P. H. Oosthuizen, Numerical investigations of buoyancy-driven natural ventilation in a

simple atrium building and its effect on the thermal comfort conditions, Applied Thermal Engineering, vol. 40, pp. 358372, Jul. 2012. [5] Y. Su, S. Riffat, Y. Lin, and N. Khan, Experimental and CFD study of ventilation flow rate of a MonodraughtTM windcatcher, Energy and Buildings, vol. 40, no. 6, pp. 11101116, 2008. [6] R. Ramponi and B. Blocken, CFD simulation of crossventilation for a generic isolated building: Impact of computational parameters, Building and Environment, vol. 53, pp. 34-48, Jul. 2012. [7] T. van Hooff, B. Blocken, L. Aanen, and B. Bronsema, A venturi-shaped roof for wind-induced natural ventilation of buildings: Wind tunnel and CFD evaluation of different design configurations, Building and Environment, vol. 46, no. 9, pp. 1797-1807, Sep. 2011. [8] K.-S. Nikas, N. Nikolopoulos, and A. Nikolopoulos, Numerical study of a naturally cross-ventilated building, Energy and Buildings, vol. 42, no. 4, pp. 422-434, Apr. 2010. [9] L. Li and C. Mak, The assessment of the performance of a windcatcher system using computational fluid dynamics, Building and Environment, vol. 42, no. 3, pp. 1135-1141, Mar. 2007. [10] S. Kirk and M. Kolokotroni, Windcatchers in modern UK buildings: experimental study, International Journal of Ventilation, vol. 3, no. 1, p. 67, 2004. [11] A. A. Elmualim and H. B. Awbi, Wind tunnel and CFD investigation of the performance of windcatcher ventilation systems, International Journal of Ventilation, vol. 1, no. 1, pp. 5364, 2002. [12] A. A. Elmualim, Effect of damper and heat source on wind catcher natural ventilation performance, Energy and Buildings, vol. 38, no. 8, pp. 939-948, Aug. 2006. [13] A. Elmualim, Verification of design calculations of a wind catcher/tower natural ventilation system with performance testing in a real building, International Journal of Ventilation, vol. 4, no. 4, pp. 393-404, 2006. [14] H. Montazeri and R. Azizian, Experimental study on natural ventilation performance of one-sided wind catcher, Building and Environment, vol. 43, no. 12, pp. 2193-2202, Dec. 2008. [15] H. Montazeri and R. Azizian, Experimental study on natural ventilation performance of a two-sided wind catcher, in Proceedings of the Institution of Mechanical Engineers, Part A: Journal of Power and Energy, 2009, pp. 387-400. [16] H. Montazeri, Experimental and numerical study on natural ventilation performance of various multi-opening wind catchers, Building and Environment, vol. 46, no. 2, pp. 370378, Feb. 2011. [17] V. Kalantar, Numerical simulation of cooling performance of wind tower (Baud-Geer) in hot and arid region, Renewable Energy, vol. 34, no. 1, pp. 246-254, Jan. 2009. [18] S. Liu and C. M. Mak, Numerical evaluation of louver configuration and ventilation strategies for the windcatcher system, Building and Environment, vol. 46, no. 8, pp. 16001616, Aug. 2011. [19] W. P. Jones and B. E. Launder, The Prediction of Laminarization with a Two-Equation Model of Turbulence,

14

International Journal of Heat and Mass Transfer, vol. 15, pp. 301-314, 1972. [20] B. E. Launder and B. I. Sharma, Application of the Energy Dissipation Model of Turbulence to the Calculation of Flow Near a Spinning Disc, Letters in Heat and Mass Transfer, vol. 1, no. 2, pp. 131-138, 1974. [21] Y. Kurazumi, T. Tsuchikawa, J. Ishii, K. Fukagawa, Y. Yamato, and N. Matsubara, Radiative and convective heat transfer coefficients of the human body in natural convection, Building and Environment, vol. 43, no. 12, pp. 2142-2153, Dec. 2008. [22] P. Ole Fanger and J. Toftum, Extension of the PMV model to non-air-conditioned buildings in warm climates, Energy and Buildings, vol. 34, no. 6, pp. 533-536, Jul. 2002.

15

You might also like