Wind driven flow through building openings

Similar documents
Published in: CleanTech for Sustainable Buildings: From Nano to Urban Scale (CISBAT), Lausanne, September 2011

Single-sided natural ventilation through a centre-pivot roof window Iqbal, Ahsan; Nielsen, Peter Vilhelm; Gunner, Amalie; Afshari, Alireza

A REVIEW OF THE CONCEPT OF DISCHARGE COEFFICIENT FOR DESIGNING NATURAL VENTILATION IN BUILDINGS

OPENING DESIGN TO IMPROVE THE NATURAL VENTILATION PERFORMANCE OF HIGH-RISE RESIDENTIAL BUILDINGS

EXPERIMENTAL AERODYNAMICS FLOW TROUGH AN ORIFICE LAB SHEET

Terrain classification and exposure factor in the 2005 National Building Code of Canada

WIND LOADS ON BALUSTRADES. Antonios W. Rofail a, Christian Mans a

Validation of Numerical Modeling of Air Infiltration through Building Entrance Doors

WIND LOADS ON BUILDINGS MWFRS (ENVELOPE PROCEDURE)

putational domain contained 8 blocks surrounding the central one. All 9 blocks had the same arrangement. Each block comprised a total of 72 residentia

WIND FLOW IN THE RECESSED CAVITIES OF A TALL

Wind Pressures on Solar Panels Mounted on Residential Homes

Numerical simulation of inter-flat pollutant dispersion routes and cross-contamination risk in high-rise residential buildings Di Mu, Naiping Gao* Ins

UNCERTAINTIES IN AIRFLOW NETWORK MODELLING TO SUPPORT NATURAL VENTILATION EARLY STAGE DESIGN

DYNAMIC WIND LOADING OF H-SHAPED TALL BUILDINGS

Airflow Prediction in Buildings for Natural Ventilation Design: Wind Tunnel Measurements and Simulation. Panagiota Karava. A Thesis.

A study of single-sided ventilation and provision of balconies in the context of high-rise residential buildings

UNCERTAINTIES IN AIRFLOW NETWORK MODELLING TO SUPPORT NATURAL VENTILATION EARLY STAGE DESIGN

Available online at ScienceDirect. Energy Procedia 78 (2015 )

Changing the geometry of the Wind Tower and its Influence on Aerodynamic Behavior and Natural Ventilation

Numerical study of the cross-ventilation of an isolated building with. different opening aspect ratios and locations for various wind.

Analysis of Uniformity and Energy Consumption in. Supermarket installed DurkeeSox Air Dispersion System. with PE Air Dispersion Model

WIND LOADS ON SOLAR PANEL SYSTEMS ATTACHED TO BUILDING ROOFS. Eleni Xypnitou. A Thesis. The Department. Building, Civil and Environmental Engineering

Wind Loads on Rooftop Solar Panel Systems: A Contribution to NBCC 2015

ADVANCES in NATURAL and APPLIED SCIENCES

Problems with Pitots. Issues with flow measurement in stacks.

COMPUTATIONAL ANALYSIS ON THE EFFECTS OF FAÇADE MODIFICATIONS ON WIND-DRIVEN NATURAL VENTILATION PERFORMANCE OF A SINGLE-CELL ROOM

Wind-induced Pressures on Patio Covers

CFD and Wind Tunnel Study of the Performance of a Multi- Directional Wind Tower with Heat Transfer Devices

CFD in Ventilation, case studies from REHVA CFD guidebook

Effect of Orifice Plate Shape on Performance Characteristics

Wind-Induced Torsional Loads on Low- and Medium-Rise Buildings

On the performance of the Stravent ventilation system in an office space Numerical and experimental investigations

THE PREDICTION OF WIND LOADS ON BUILDING ATTACHMENTS

ASSESSMENT OF AIR CHANGE RATE AND CONTRIBUTION RATIO IN IDEALIZED URBAN CANOPY LAYERS BY TRACER GAS SIMULATIONS

Impact of airtightness on the heat demand of passive houses in central European climate

Local and overall wind pressure and force coefficients for solar panels

Prediction of wind pressure coefficients in building energy simulation using machine learning

Analysis of Wind Loads on Buildings and Signs: A Computer Program Based on ASCE 7

SIMULATION OF THE PERFORMANCE OF A HYBRID VENTILATION SYSTEM IN DIFFERENT CLIMATES. Technicka 2896/2, Brno, Czech Republic

Energy Consumption and Indoor Environment Predicted by a Combination of Computational Fluid Dynamics and Building Energy Performance Simulation

Numerical Investigation of the Integration of Heat Transfer Devices into Wind Towers

MODELLING THE URBAN MICROCLIMATE AND ITS INFLUENCE ON BUILDING ENERGY DEMANDS OF AN URBAN NEIGHBOURHOOD

MODELLING BUOYANCY INDUCED FLOWS OF PASSIVE COOLING SYSTEMS Pedro Correia da Silva 1, Vítor Leal 1 and J. Correia da Silva 2

APPLYING COMPUTATIONAL FLUID DYNAMICS TO ARCHITEC- TURAL DESIGN DEVELOPMENT

ANALYSIS AND MITIGATION OF THERMAL EFFECTS ON A LARGE FACADE-MOUNTED PV ARRAY

NATURAL VENTILATION NETWORK DESIGN OF A BUILDING

Effect of Building Shape And Neighborhood Design On Solar Potential And Energy Performance

A model for calculating single-sided natural ventilation rate in an urban residential apartment

Experimental Research on the Heat Transfer and Flow Performance of a Composite Heat Sink

NUMERICAL STUDY ON FILM COOLING AND CONVECTIVE HEAT TRANSFER CHARACTERISTICS IN THE CUTBACK REGION OF TURBINE BLADE TRAILING EDGE

Multi-scale methodology to assess wind loads on building louvers

Infiltration simulation in a detached house empirical model validation

Numerical Modeling of Buoyancy-driven Natural Ventilation in a Simple Three Storey Atrium Building

A STUDY OF CROSS CONTAMINATION OF IN-SUITE VENTILATION SYSTEMS USED IN MULTI-UNIT RESIDENTIAL BUILDINGS

LITERATURE REVIEW OF NATURAL VENTILATION Pedro Romero

MULTIPLE-INLET BIPV/T: MODELLING AND DESIGN INCLUDING WIND EFFECTS. Quebec, Canada

MODELLING THE ENERGY PERFORMANCE OF NIGHT-TIME VENTILATION USING THE QUASI-STEADY STATE CALCULATION METHOD

Effect of Rotation Speed of a Rotary Thermal Wheel on Ventilation Supply Rates of Wind Tower System

INTERNATIONAL JOURNAL OF CIVIL AND STRUCTURAL ENGINEERING Volume 4, No 3, 2014

Assessment of the natural air ventilation of buildings in urban area with the CFD tool UrbaWind. Dr Stéphane SANQUER Meteodyn

CFD/FEM Based Analysis Framework for Wind Effects on Tall Buildings in Urban Areas

Proceedings of Clima 2007 WellBeing Indoors

Urban Design to Reduce building energy use: Using CFD to inform better placement of buildings.

THERMAL ANALYSIS OF A FACADE-MOUNTED PV ARRAY

CFD INVESTIGATION OF AIRFLOW AROUND CONIC TENSILE MEMBRANE STRUCTURES

Riccardo Buccolieri 1, Mats Sandberg 2

Evaluation of single-sided natural ventilation using a simplified and fair calculation method

For all those with the ability to influence the design, layout or indoor air quality of buildings

FAST EVALUATION OF SUSTAINABLE HEATING AND COOLING STRATEGIES FOR SOLAR HOMES WITH INTEGRATED ENERGY AND CFD MODELING

A MODEL-BASED METHOD FOR THE INTEGRATION OF NATURAL VENTILATION IN INDOOR CLIMATE SYSTEMS OPERATION

Numerical Modelling of Air Distribution in the Natatorium Supported by the Experiment

Numerical study of the cross-ventilation of an isolated building with different opening aspect ratios and locations for various wind directions

H POLLUTANT DISPERSION OVER TWO-DIMENSIONAL IDEALIZED STREET CANYONS: A LARGE-EDDY SIMULATION APPROACH. Colman C.C. Wong and Chun-Ho Liu*

Wind loads at solar and photovoltaic modules for large plants

u and Yoshie (2013) investigate the effect of building geometry and they revealed that the ventilation efficiency strongly depended on the level of bu

Javanese House s Roof (Joglo) with the Opening as a Cooling Energy Provider

BUILDING INTEGRATED VENTILATION SYSTEMS MODELLING AND DESIGN CHALLENGES

An Experimental Investigation into the Pressure Leakage Relationship of some failed water pipes

A Sensitivity Analysis on Mixing Energy Loss in Air-Conditioned Rooms by Using CFD

Vulnerability modelling of metal-clad industrial buildings to extreme wind loading

Sensitivity Analysis of Personal Exposure Assessment Using a Computer Simulated Person Brohus, Henrik; Jensen, H. K.

EXAMINATION OF TRAFFIC POLLUTION DISTRIBUTION IN A STREET CANYON USING THE NANTES 99 EXPERIMENTAL DATA AND COMPARISON WITH MODEL RESULTS

Investigation of the effect of balconies on natural ventilation of dwellings in high-rise residential buildings in subtropical climate

Laboratory Testing of Safety Relief Valves

PERFORMANCE OF AN AUXILIARY NATURAL VENTILATION SYSTEM

Application of CFD Predictions to Quantify Thermal Comfort for Indoor Environments

Steven J. Emmerich and W. Stuart Dols National Institute of Standards and Technology Gaithersburg, MD, USA

THERMAL ENVIRONMENT OF OUTDOOR UNITS OF VRV SYSTEM IN HIGH- RISE BUILDING. Gang Wang, Yafeng Hu, and Songtao Hu

PRELIMINARY RESULTS FROM FIELD EXPERIMENTAL STUDY OF RAIN LOAD AND PENETRATION INTO WOOD-FRAME WALL SYSTEMS AT WINDOW SILL DEFECTS

Torsional and Shear Wind Loads on Flat-Roofed Buildings

THERMOELECTRIC EFFECTS OF SIZE OF MICROCHANNELS ON AN INTERNALLY COOLED LI-ION BATTERY CELL

A numerical investigation of the influence of wind on multiple short natural draft dry cooling towers

SMART CONTROL FOR NATURAL VENTILATION HOUSE

HY-12 User Manual. Aquaveo. Contents

USE OF MULTI-ZONE AIR FLOW SIMULATIONS TO EVALUATE A HYBRID VENTILATION SYSTEM. Åke Blomsterberg 1, Tomas Johansson 1.

PARAMETRIC CFD WIND FORCE ANALYSIS ON A RESIDENTIAL ROOF MOUNTED PHOTOVOLTAIC PANEL

TREES IN URBAN STREET CANYONS AND THEIR IMPACT ON THE DISPERSION OF AUTOMOBILE EXHAUSTS. Christof Gromke*, Bodo Ruck

Numerical Investigation of the Air Flow in a Novel PV-Trombe Wall Based on CFD Method

Transcription:

International Conference Passive and Low Energy Cooling 427 for the Built Environment, May 25, Santorini, Greece Wind driven flow through building openings P. Karava, T. Stathopoulos and A.K. Athienitis Department of Building, Civil and Environmental Engineering, Concordia University, Montreal, Quebec, Canada ABSTRACT This paper presents internal pressure coefficients and discharge coefficients in a building with wind-driven cross-ventilation caused by sliding window openings on two adjacent walls. The study found that both coefficients vary considerably with the opening area and the inlet to outlet ratio. 1. INTRODUCTION Field experiments carried out in an urban canyon have shown that appreciable ventilation rates can be obtained with natural ventilation, especially when cross ventilation with two or more windows is used (Niachou et al., 25). However, the modeling of large openings (windows partly open) remains a significant source of uncertainty in prediction of ventilation performance. In fact, comparisons of simplified empirical methodologies with respect to airflow prediction show significant variations when different equations or solution methods are used. The most common equation describing the airflow through an opening is the orifice equation, which is based on Bernoulli s assumption for steady incompressible flow, and requires that the discharge coefficient, C D, and the internal pressure coefficient,, are both known. For typical low-rise residential buildings, the airflow through ventilation openings (i.e. openable windows) is mainly wind-driven, especially during the summer. For wind-driven ventilation, the discharge coefficient for an inlet can be determined by: CD,inlet Q u = = (1) A V Cpw Cpin V Cpw Cpin in which Q: airflow through the opening, u: velocity in the opening, A: opening area, V: reference wind speed at the building s height, C pw : pressure coefficient on windward façade, C pin : internal pressure coefficient, determined for a building with two openings by Equations (2) and (3) for large openings (windows, doors turbulent flow) and small openings (cracks laminar flow) respectively: C 2 Cp2 +α Cp1 pin= 2 (2) 1+α Cp2 +β Cp1 Cpin = (3) 1+β where C p1 : external pressure coefficient in opening 1, C p2 : external pressure coefficient in opening 2, α=a 1 /A 2 and β=¼ for uniform distribution of cracks. Generally, C pin depends on the external pressure distribution, terrain, shape, area and distribution of openings on the façade. This paper presents the results of a series of experiments carried out in a Boundary Layer Wind Tunnel (BLWT) for the evaluation of the internal pressure coefficient in a building with cross-ventilation. Discharge coefficient values for different opening configurations with different inlet to outlet ratio are presented. 2. DESCRIPTION OF THE MODEL AND EXPERIMENTAL SET UP A 1:12 gable roof-sloped building model of rectangular plan view 15.3 9.8 cm was tested with an eave height of 3 cm. This corresponds to a building 61 39 m and 12 m high accord-

428 International Conference Passive and Low Energy Cooling for the Built Environment, May 25, Santorini, Greece ing to the 1:4 geometric scale in the BLWT of Concordia University. The model provides variable side-wall and windward wall openings and background leakage of and.5%. The background leakage was achieved by a series of holes which could be left open or closed. Five pairs of closely-located internal and external pressure taps have been selected for the measurements as indicated in Figure 1. Plastic tubes connect each tap with a Honeywell 163 PC pressure transducer. More details on the building model can be found in Wu et al. (1998). A standard wind profile over open country terrain was simulated with a power-law exponent equal to.15 and turbulence intensity at the eave height equal to 22%. In this study only simple rectangular openings of the same height (1.7 cm) were considered. Differences in opening areas were induced by different opening width (sliding windows). The wall thickness is equal to 3 mm. A singlesided ventilation configuration was considered first for comparison proposes. Experiments were performed for and.5% background leakage. This was followed by cross-ventilation experiments with.5% background leakage. Configurations with equal inlet and outlet opening area were tested as well as configurations with inlet to outlet ratio smaller or larger than 1. The two rectangular openings were located in the middle of long wall (windward wall) and short wall (side wall). Measurements were carried out for a wind angle, θ, equal to. Figure 1 shows the different opening configurations considered in this study. 3. RESULTS AND DISCUSSION 3.1 Internal pressure coefficient The measured mean values of the external pressure coefficients (reference height=building height) are.67 for the windward wall, -.36 for the side wall, -.25 for the leeward wall and -.4 and -.74 on taps 7 and 9 of the roof. The measured mean internal pressure coefficient for.5% background leakage (without openings) is -.36 which is slightly different from the theoretical value, -.23, obtained by Equation (3) using C p1 as the representative of the positive external pressure and C p2 as the area-averaged pressure on the rest of the building envelope. Figure 1: Exploded view of building model with pressure tap locations and opening configurations. 5 1 15 2 25 3.8.6.4.2 Opening porosity (%) 2 4 6 8 1 12 14 -.2 Opening area (cm 2 BLWT, % leakage ) -.4 BLWT,.5% leakage -.6 Figure 2: Internal pressure coefficients for single-sided ventilation and different opening porosity. The impact of a windward wall opening (single-sided ventilation) on internal pressure was investigated for and.5% background leakage and results are presented in Figure 2 for different opening area or opening porosity (A opening/a wall ). The opening is located in the middle of the long façade. The internal pressure was measured at different internal taps and it was found to be uniform, as also previously reported by Wu et al. (1998). The experimental data is compared with the values obtained by Equation (2) for % background leakage. For the case of a single opening, reduces to =Cp 1, which is equal to.67.

International Conference Passive and Low Energy Cooling 429 for the Built Environment, May 25, Santorini, Greece Figure 2 shows good agreement between the experimental results and the theoretical values for % leakage. In addition, data obtained for.5% background leakage shows a similar trend with that observed in previous studies, e.g. Wu et al. (1998). However, cannot be used in this case due to the undetermined character of the flow. The impact of a windward and a side-wall opening (cross ventilation) of the same area (A 1 =A 2 ) on internal pressure was investigated for.5% background leakage. The openings were located in the middle of the long and short walls see Figure 1. Measurements were carried out for opening area up to 1.2 cm 2 (or 22 % opening porosity); which is typically the range in naturally ventilated houses. The external pressure distribution was monitored and found not affected by the presence of openings on the façade (sealed body assumption). The internal pressure was measured on taps 2, 4, 6, 8 and 1. Figure 3 presents the mean value and standard deviation of the internal pressure coefficient as a function of the opening area. The values by using are also presented. For A 1 /A 2 =1, reduces to =(Cp 1 +Cp 2 )/2. Note that the background leakage is not considered when using Equation (2). The experimental results show that the average increases with the increase of the opening area, although provides a constant value. It was found that the internal pressure is not uniform for opening area larger than about 5 cm 2 (which corresponds to approximately 1% opening porosity), resulting in a increase of the standard deviation of the internal pressure coefficient. This is probably due to a virtual flow tube that connects the inlet and.8.6.4.2 2 4 6 8 1 12 -.2 Opening area (cm 2 ) -.4 BLWT, A1 = A2 -.6 Figure 3: Internal pressure coefficients for cross- ventilation, A 1 =A 2 (α=1) and.5% background leakage. the outlet see also Murakami et al. (1991), Sawachi et al. (24). In fact, small values were measured in taps 2 and 1 that are in the flow tube, while high values were recorded in taps 4, 6 and 8 that are not in the flow tube. These variations should be considered in the selection of the measurement points. The nonuniformity of distribution in the room was not observed in the case of single-sided ventilation and it is not predicted by the theory Equation (2). Variation in values along the flow tube have been reported by Sawachi et al. (24). In the same study, the average in a cross ventilated room with two rectangular openings with 9% inlet and outlet opening porosity was about.52, while in the study by Murakami et al. (1991) varied from -.33 to.18 for different configurations and opening areas ranging from 5 to 36 cm 2. The internal pressure coefficient was also investigated for inlet opening area equal to 2.6 cm 2 (A 1 =2.6 cm 2 ),.5% background leakage and different inlet to outlet ratio (outlet opening area, A 2, varied from.9 to 12 cm 2 ). The experiments were repeated for inlet opening area, A 1, equal to 5.2 cm 2. The internal pressure was measured on taps 2, 4, 6, 8 and 1. The mean and standard deviation of the internal pressure coefficient as a function of inlet to outlet ratio (A 1 /A 2 ) for A 1 =2.6 cm 2 is shown in Figure 4 and for A 1 =5.2 cm 2 in Figure 5. The results are compared with data from Murakami et al (1991) for the case of a side-wall outlet (model 1) and a leeward wall outlet (model 2). The values by using are also presented. The experimental results indicate that the average increases with the increase of inlet to outlet ratio. For A 1 /A 2 <1 (i.e. large outlet area, A 2 ) the standard deviation of is higher due to differences among the various internal pressure taps; these differences are even higher for larger inlet area, A 1 =5.2 cm 2. This internal pressure coefficient variation is again due to the flow tube connecting the inlet and the outlet. The values predicted by are overestimated compared to the experimental data, particularly for higher inlet to outlet ratios. This might be due to the impact of the background porosity that is more important for small opening areas and high A 1 /A 2 ratios. Comparison of Figures 4 and 5 shows that there are no substantial differences among average values for

43 International Conference Passive and Low Energy Cooling for the Built Environment, May 25, Santorini, Greece the same A 1 /A 2 ratio but different A 1 and A 2, as opposed to values for configurations with A 1 =A 2 (Figure 3). However, the internal pressure coefficient varies considerably (from -.26 to.47) for configurations with A 1 /A 2 >1 or A 1 /A 2 <1 compared to configurations with A 1 /A 2 =1 (from -.5 to.19). This might be important to be considered in natural ventilation design. 3.2 Inlet discharge coefficient 3.2.1 Theoretical background The discharge coefficient (Equation 1) is determined by applying the Bernoulli equation on a horizontal stream line between a point in front of the building opening with stagnant air and pressure and the vena contracta (minimum cross section area of the flow with parallel stream lines, uniform velocity and static pressure equal to the surrounding air pressure) see also Andersen (1996) and Karava et al. (24). Hence, in case an opening separates two regions with still air boundary conditions, the discharge coefficient can be assumed to be constant ranging from.6 to 1, depending primarily on the shape (geometry) of the opening. Bernoulli equation assumes two points that are not affected by the opening and the pressure if uniform. Thus selection of the measurement point should not be a problem. For wind-driven cross-ventilation the velocity and pressure fields at the inlet are unsteady, creating difficulties with the selection of C D and, more fundamentally, its definition (Etheridge, 24). Experimental results confirm the previous statement, particularly in the case of cross-ventilation with large opening areas (see Figures 3, 4 and 5). Thus, care should be taken in the selection of the measurement point. For wind-driven cross-ventilation, the discharge coefficient depends on the geometry of the opening but also the external (surrounding and building) and the internal flow field (background leakage, leeward/side wall openings). In fact, for large opening area, there is significant air movement in the room due to the flow tube and the flow might no longer be considered as pressure driven. Therefore, discharge coefficient values might be outside the standard range i.e. from.6 to 1. 3.2.2 Experimental results In order to determine the inlet discharge coefficient, the velocity ratio u/v - see Equation (1) - was evaluated using a hotfilm anemometer. The results are presented in Table 1 along with data from other similar wind tunnel studies reported in the literature. The ratio u/v is about.6-.63 in the present study and it varies between.45 and.83 in other literature sources. A more precise method to evaluate this ratio might be to consider velocity field measurements (PIV technique), as opposed to point measurement techniques. In the present study, the velocity in the opening is assumed equal to the velocity just before the opening, i.e. for two inlet openings with dif- Table 1: Velocity ratios (u/v) for different studies. Reference u/v U ref (m/s) Present study.6-.63 7.2 Hu (25).45 7. Etheridge (24).6 4. Sawachi et al. (24).5 3. Murakami et al. (1991).62 1. Not reported.8.6.4.2 -.2 1 2 3 4 5 6 7 8 A 1 /A 2 BLWT, A1 = 2.6 sq. cm -.4 Murakami et al. (1991), model 1 Murakami et al. (1991), model 2 -.6 Figure 4: Internal pressure coefficients for cross- ventilation and different inlet to outlet ratios (A 1 =2.6 cm 2 )..8.6.4.2 1 2 3 4 5 6 7 8 -.2 A 1 /A BLWT, A1 = 5.2 sq. cm 2 -.4 Murakami et al. (1991), model 1 Murakami et al. (1991), model 2 -.6 Figure 5: Internal pressure coefficients for cross- ventilation and different inlet to outlet ratios (A 1 =5.2 cm 2 ).

International Conference Passive and Low Energy Cooling 431 for the Built Environment, May 25, Santorini, Greece ferent areas the velocity ratio (u/v) is assumed to be the same. Consequently, the difference in the discharge coefficient between the two openings is due to variation with the opening area. Velocity profile measurements carried out by Jensen et al. (22) show that for opening porosity from 2 to 2 % (porosity range considered in the present study), the ratio of the velocity in the opening and the velocity just before the opening is about.9-1. (except for measurement points close to the edge). This confirms that the assumption made in the present study is reasonable. Figure 7 shows the variation of the inlet discharge coefficient with the opening area for cross-ventilation with two rectangular openings of the same area (A1=A 2 ) located on the windward and side wall of the building model for.5% background leakage. Wind tunnel data from Murakami et al. (1991) (for crossventilation with openings located in the long façade of the building) and Jensen et al. (22) are also included. Most previous studies have found that the discharge coefficient increases with the increase of the opening area. However, Sawachi et al. (24) has reported that C D is not affected by the opening area. Heiselberg et al. (22) found that C D might decrease, increase or remain almost constant depending on the configuration, regardless of opening area. Karava et al. (24) reviewed the current literature and concluded that no clear trend can be established. The results of the present study show that the inlet discharge coefficient varies from.74 to.9 for opening area from.1 to 1 cm 2 (corresponding to 2-2% inlet opening porosity) and u/v=.63 and from.6 to.71 for Inlet C D 1.2 1.8.6.4.2 C D = 1 C D =.65 BLWT, u/v =.63 BLWT, u/v =.5 Jensen et al. (22) Murakami et al. (1991), u/v =.64-1 1 2 3 4 Opening area (cm 2 ) Figure 7: Inlet C D as a function of the opening area for equal inlet to outlet ratio. u/v=.5. Different C D values are observed in the various studies especially for u/v=.63 due to different external flow conditions, opening and building configurations considered. A.5% background leakage was considered in the present study as opposed to impermeable models used in other studies. Hence, the results of a study can only be used within the limits of their applicability. For the opening and building configurations considered in Figure 7, the discharge coefficient is higher than.65, which is the typical value given in textbooks, e.g. Etheridge and Sandberg (1996). For u/v=.5 the BLWT results of the present study are in good agreement with the other literature sources. Therefore, C D =.65 might be a good approximation for velocity ratio u/v=.5. However, further experimental work is required considering different building and opening configurations as well as more precise measurement techniques such as PIV before any generalization is to be made. Figure 8 shows the variation of inlet discharge coefficient with the inlet to outlet ratio (A 1 /A 2 ) for A 1 =2.6 cm 2. Data from Murakami et al. (1991) were also considered for the case of cross ventilation with a windward wall inlet and a side wall outlet. Generally, the discharge coefficient varies from.65 to 1.15 for u/v equal to.63 and from.52 to.9 for u/v equal to.5. For A 1 /A 2 <1 there is smaller variation of C D compared to that for A 1 /A 2 >1. A similar observation was made by Sandberg (24) for the catchment area Ac, defined as Q/Uo for relative large opening area. Non standard discharge coefficients (i.e C D >1) are observed for A 1 /A 2 >>1. This can be justified by the unsteady external and internal flow field in the case of Inlet C D 1.2 1.8.6.4.2 C D = 1 C D =.65 u/v =.63 u/v =.5 Murakami et al. (1991), u/v =.64.1 1 1 A1/A2 Figure 8: Inlet C D as a function of inlet to outlet ratio for A 1 =2.6 cm 2.

432 International Conference Passive and Low Energy Cooling for the Built Environment, May 25, Santorini, Greece cross ventilation particularly for large opening area, as previously discussed. 4. CONCLUSIONS The internal pressure coefficient varies considerably for cross-ventilation configurations with unequal inlets and outlets. This is particularly important for natural ventilation design. The inlet discharge coefficient varies with the opening area and inlet to outlet ratio, particularly for A 1 /A 2 >1. Wu H., Stathopoulos T., Saathoff P. (1998). Windinduced internal pressures revisited: low-rise buildings, Structural Engineers World Congress, San Francisco, CA, USA. REFERENCES Andersen, K.T., 1996. Inlet and outlet coefficients: a theoretical analysis, Proceedings of Roomvent 1996, 1, pp. 379-39. Etheridge, D. and M. Sandberg, 1996. Building Ventilation: Theory and Measurement, John Willey & Sons, London. Etheridge, D.W., 24. Natural ventilation through large openings measurements at model scale and envelope theory. International Journal of Ventilation, 2, (4), pp. 325-342. Heiselberg, P. E. Bjorn and P.V. Nielsen, 22. Impact of open windows on room air flow and thermal comfort, International Journal of Ventilation, 1, (2). Hu, H.C., 25. Computer simulations of pedestrian wind environment around buildings and numerical studies of wind effects on cross-ventilation. Wind Effects Bulletin, 3, Feb., Tokyo Polytechnic University. Jensen, J.T, M. Sandberg, P. Heiselberg and P.V. Nielsen, 22. Wind driven cross-flow analyzed as a catchment problem and as a pressure driven flow, International Journal of Ventilation, HybVent-Hybrid Ventilation Special Edition, 1, pp. 88-11. Karava, P., T. Stathopoulos and A.K. Athienitis, 24. Wind-driven flow through openings: A Review of discharge coefficients, International Journal of Ventilation, 3(3), pp. 255-266 Liu, H., 1991. Wind Engineering A Handbook for structural engineers, Prentice-Hall, New Jersey. Murakami, S., S. Kato, S. Akabashi, K. Mizutani and Y.- D. Kim, 1991. Wind tunnel test on velocity-pressure field of cross-ventilation with open windows, ASH- RAE Transactions, 97, (1), pp. 525-538. Niachou, K., S. Hassid, M. Santamouris and I. Livada, 25. Comparative monitoring of natural, hybrid and mechanical ventilation systems in urban canyons, Energy and Buildings, 37, pp. 53-513. Sandberg, M., 24. An alternative View on Theory of Cross-Ventilation, International Journal of Ventilation, 2(4), pp. 4-418. Sawachi, T., K. Narita, N. Kiyota, H. Seto, S. Nishizawa and Y. Ishikawa, 24. Wind pressure and airflow in a full-scale building model under cross ventilation, International Journal of Ventilation, 2, (4), pp.343-357.