• No results found

Overview of pressure coefficient data in building energy simulation and airflow network programs

N/A
N/A
Protected

Academic year: 2021

Share "Overview of pressure coefficient data in building energy simulation and airflow network programs"

Copied!
31
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Overview of pressure coefficient data in building energy

simulation and airflow network programs

Citation for published version (APA):

Costola, D., Blocken, B. J. E., & Hensen, J. L. M. (2009). Overview of pressure coefficient data in building energy simulation and airflow network programs. Building and Environment, 44(10), 2027-2036.

https://doi.org/10.1016/j.buildenv.2009.02.006

DOI:

10.1016/j.buildenv.2009.02.006 Document status and date: Published: 01/01/2009

Document Version:

Accepted manuscript including changes made at the peer-review stage

Please check the document version of this publication:

• A submitted manuscript is the version of the article upon submission and before peer-review. There can be important differences between the submitted version and the official published version of record. People interested in the research are advised to contact the author for the final version of the publication, or visit the DOI to the publisher's website.

• The final author version and the galley proof are versions of the publication after peer review.

• The final published version features the final layout of the paper including the volume, issue and page numbers.

Link to publication

General rights

Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights. • Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain

• You may freely distribute the URL identifying the publication in the public portal.

If the publication is distributed under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license above, please follow below link for the End User Agreement:

www.tue.nl/taverne Take down policy

If you believe that this document breaches copyright please contact us at: openaccess@tue.nl

(2)

* Corresponding author: D. Costola

Building Physics and Systems, Eindhoven University of Technology, P.O. box 513, 5600 MB Eindhoven, the

Overview of pressure coefficient

1

data in building energy simulation programs

2

3

D. Cóstola

*

, B. Blocken, J.L.M. Hensen

4

5

Building Physics and Systems, Eindhoven University of Technology, the Netherlands

6 7 8

Abstract 9

Wind pressure coefficients (Cp) are influenced by a wide range of parameters, including building 10

geometry, surrounding terrain topography, facade detailing, position on the facade and wind direction. As 11

it is practically impossible to take into account the full complexity of pressure coefficient variation, 12

Building Energy Simulation (BES) programs generally incorporate it in a simplified way. This paper 13

provides an overview of pressure coefficient data and the extent to which they are currently implemented 14

in BES programs. A distinction is made between primary sources of Cp data, such as full scale 15

measurements, reduced-scale measurements in wind tunnels and computational fluid dynamics (CFD) 16

simulations, and secondary sources, such as databases and analytical models. The comparison between 17

data from secondary sources implemented in BES shows that the Cp values are quite different depending 18

on the source adopted. The two main parameters causing these differences are the position on the facade 19

and the sheltering effects. The comparison of Cp data for sheltered buildings shows the largest 20

differences, and data from different sources present different trends. The paper concludes that 21

quantification of the uncertainty related to such data sources is required to guide future improvements in 22

Cp implementation in BES programs. 23

24 25

Keywords: wind pressure coefficient, building energy simulation (BES), airflow network (AFN); 26

ventilation; infiltration; model intercomparison; building envelope 27

28

Cóstola, D., Blocken, B. & Hensen, J.L.M. (2009).

Overview of pressure coefficient data in building energy simulation and airflow network programs. Building and Environment, 44(10), 2027-2036

(3)

1 Introduction 1

Air infiltration and ventilation have a profound influence on both the internal environment and the 2

energy needs of buildings [1]. Air flow through the building envelope is also an important factor 3

influencing building heat loss [2]. 4

Wind is an important driving force for infiltration and ventilation. Wind pressure is therefore an 5

important boundary condition for a wide range of models, from building component heat, air and 6

moisture (HAM) transfer models to coupled airflow network (AFN) and building energy simulation 7

(BES) programs [3-6]. 8

Wind pressure on the building envelope is usually expressed by pressure coefficients (Cp), which are 9 defined as follows: 10 x 0 P d

P

P

C

P

; 2 h d

U

P

2



(1) 11

where Px is the static pressure in a given point at the building facade (Pa), P0 is the static reference 12

pressure (Pa), Pd is the dynamic pressure in the upstream undisturbed flow (Pa),  is the air density 13

(kg/m3) and Uh is the wind speed at building height h in the upstream undisturbed flow (m/s). 14

The impact of Cp in BES results was studied using sensitivity analysis, identifying Cp as one of the 15

main sources of uncertainty in BES-AFN models [7-8]. The reasons are the high uncertainty associated 16

with Cp values, and the fact that several performance indicators, e.g. energy consumption, thermal 17

comfort and mould growth, are often very sensitive to the air change rate, which depends on Cp. 18

Hensen [9] described the difficulty to perform an accurate evaluation of Cp. Further studies, especially 19

on the development of analytical models for Cp prediction [10,11], have not been able to overcome the 20

difficulties in obtaining reliable Cp values without using expensive wind tunnel experiments. In this sense, 21

Tuomaala [12] stated that ―there is no reliable and effective method for evaluating the value of wind 22

pressure coefficient for complex cases.‖ 23

The difficulties in assessing reliable Cp data for BES can be explained by the wide range of 24

influencing parameters, including building geometry, facade detailing (e.g. external shading devices, 25

balconies), position on the facade, sheltering elements (e.g. buildings, trees), wind direction and 26

turbulence intensity,. 27

(4)

To the knowledge of the authors, there is no overview of pressure coefficient data for indoor air flow 1

studies in BES. The purpose of the present paper therefore is to provide such information together with a 2

comparison of data from different data sources. 3

This overview can be useful for the development and validation of AFN, BES and HAM models, for 4

ventilation, infiltration and indoor air quality studies, for documentation for building certification 5

programs and for the development of new models for Cp prediction. 6

The paper is structured as follows. Section 2 describes the BES and AFN programs that are included 7

in this overview, as well as the method for their selection. The sources of Cp data are classified in primary 8

and secondary sources. Section 3 provides a brief overview of primary sources: full scale measurements, 9

wind tunnel tests and CFD simulations. Primary sources provide data for a specific building and take into 10

account most of the parameters that influence the Cp data. Data from primary sources is, in general, more 11

accurate than secondary sources data, but the cost to obtain them is also higher. Sections 4 to 6 describe 12

the secondary sources in detail. The term secondary sources is used in this paper to describe those sources 13

that were generated based on primary sources. A large number of secondary Cp sources exist, which have 14

different aims and characteristics. Databases of Cp values are the most common secondary source. Those 15

databases, which are collection of Cp data, are present in several books and standards, and they provide 16

data for a limited set of generic building configurations. Another secondary source are the analytical 17

models, where primary data is used to derive general equations to predict Cp. Section 7 presents 18

comparisons of the values from each data source for the simple case of a cubic building, to demonstrate 19

the magnitude of the differences between them. Finally, the main conclusions are provided in section 8. 20

2 BES and AFN programs 21

This section presents the programs selected to take part in the present overview. It is not the intention 22

of this selection to be complete, rather the purpose is to obtain a reliable sample of programs representing 23

the state of the art in commercial and academic BES and AFN modelling. 24

The selection of BES programs was made based on the recent work of Crawley et al. [13]. This paper 25

results from the collaborative work of many research groups dealing with BES, and contrasts capabilities 26

of 20 BES programs. From those 20 programs, 7 included a fully implemented and coupled AFN [13], 27

which requires Cp data. 28

(5)

Stand-alone AFN tools also represent an important reference for the present paper. Firstly because 1

some of them are coupled with BES tools, e.g. Trnsys and COMIS [14], and secondly because they may 2

include the most advanced features which will then be implemented in BES tools. The most recent review 3

of AFN tools was published in 1999, comparing 14 AFN programs [15]. From those, 3 programs claimed 4

to have some data on Cp: AIOLOS, COMIS and Nitecool. From this list, AIOLOS and COMIS are 5

included in this research. CONTAM is also mentioned in [15], but at that time the version ―Contam 96‖ 6

did not have any Cp data. The present version (CONTAMW 2.4b) does include Cp data, and is therefore 7

also included in the present overview. 8

Table 1 summarizes information about the selected programs, and the documentation used as 9

reference for this paper. In all the cases, the ―Help‖ available in the program interface was also considered 10

as source of information. 11

This set is certainly small compared to the amount of existing BES tools, but it is considered to 12

provide a representative and general overview about the topic. 13

3 Primary sources 14

Primary sources are considered to be the most reliable Cp data sources. All the tools listed in Table 1 15

allow the input of user defined Cp data, which is most often obtained from primary sources. 16

In this section, a brief description of the main primary sources is provided, focusing on their 17

advantages and disadvantages. 18

3.1 Full scale measurements 19

On site full scale measurements at real building facades provide the most representative description of 20

the pressure at the building facade. In those measurements, there is no need to reproduce boundary 21

conditions, no scaling issues, and no physical models to be adopted. However, full scale measurements 22

are complex and expensive, and are therefore mainly used for validation purposes. 23

Early full scale experiments, e.g [25], used sensors with high uncertainty for the pressure 24

measurements, such as manometers, and for the wind speed, such as cup anemometers. 25

More recent experiments using ultrasonic anemometer and pressure transducers provide a large 26

amount of high quality data about the pressure at the building facade. Raw data from full scale 27

(6)

measurements is some time available at the web site of some research centers, e.g 1

http://www.wind.ttu.edu/Research/FullScale/WERFL_4th.php. 2

In many of the past experiments, the building is relatively unsheltered, so the approaching wind flow 3

can be easily measured [26-28]. The definition of proper boundary conditions is a main constraint in full 4

scale experiments. Data produced by several measurement campaigns can not be used for validation 5

purposes due to the lack of data on the approaching wind flow [29]. In the urban environment, where 6

many nearby obstructions are present, defining a good reference measurement location becomes a main 7

challenge, and this type of experiments is less commonly found in the literature. 8

Despite the problems associated with reference measurements, on site full scale experiments are 9

always useful to gain insight in the pressure variation on the building facade in space and time, which is a 10

main concern in the structural design. However, for ventilation and infiltration studies, the effects of 11

turbulence on the pressure on the building facade can be often neglected [30;31]. 12

It is commonly assumed that the wind pressure coefficient is independent of the wind speed. This is 13

true when the flow around the building is Reynolds number independent and thermal processes do not 14

significantly influence the air flow and pressure distribution around the building. For bluff bodies with 15

sharp edges, the first assumption is generally true even for low velocities because of the high Re numbers 16

in building aerodynamics. The second assumption is not necessarily true for low wind speeds, because the 17

mean wind speed as well as turbulence can be significantly influenced by solar radiation [32]. This point 18

is not relevant for in structural aerodynamics, because there the focus is generally on strong winds under 19

neutral atmospheric stability; however infiltration and ventilation studies have to deal with all the range of 20

wind speeds. Full scale experiments conducted to assess the uncertainty in Cp prediction for natural 21

ventilation indicated very scattered values for low wind speeds [33]. The solution adopted was to discard 22

from the analysis all values where the reference wind speed was lower than 4 m/s [33]. This solution 23

clearly compromises the applicability of the results for a large range of situations with low wind speed. 24

The uncertainty in the measurements is rarely the object of detailed analysis, as prescribed by ISO 25

[34]. Cp is a derived quantity; therefore it demands measurements of several quantities: pressure on the 26

facade, reference velocity, and also air temperature and atmospheric pressure to obtain the air density. 27

Each of these parameters will contribute to the resulting Cp uncertainty, not only due to the sensor 28

uncertainty, but also due to other sources, such as the measurement protocol. The combined uncertainty 29

(7)

of full scale measurements is important because it provides the limit of accuracy when these data is used 1

for validation purposes. In the best scenario, the uncertainty on the method under validation can be 2

described as equal to the uncertainty on the reference value, but never smaller. 3

It can be concluded that full scale experiments are the primary data source that provide the most 4

representative information; however the use of these experiments is restricted to research and validation 5

purposes. Full scale experiments for urban environment and for low wind speeds still present limitations 6

to be overcome by future studies, and the uncertainty of the measurements demands further attention. 7

3.2 Wind tunnel measurements 8

Wind tunnel experiments are considered the most reliable source of pressure data for buildings in the 9

design phase. Structural engineering uses custom wind tunnel experiments to assess the wind load for a 10

specific building, considering its geometry, surrounding and wind profile in the site. The use of wind 11

tunnel measurements to provide BES input data however is limited due to cost, time and know-how 12

involved in this type of experiments. 13

In the first half of the 20th century, knowledge on wind flow around buildings was mainly established 14

using wind tunnel experiments [35]. In this early stage, the use of laminar uniform flows was common, 15

and the deficiencies of this technique were not identified for many years. Later on, the comparison 16

between wind tunnel results and full scale measurements highlighted the importance of atmospheric 17

boundary layer turbulence and it incited the development of the boundary layer wind tunnels [25]. 18

Nowadays, wind tunnel modelling techniques for mean pressure around buildings are widely available 19

[29]. However, wind tunnel experiments, as any laboratory measurements, demand special care. A 20

common exercise carried out by twelve institutions compared the wind tunnel results for the simple case 21

of an isolated cube, for 3 wind directions [36]. Based on the standard deviation published in the paper, 22

assuming a normal distribution and a confidence interval of 95%, it can be say that the overall variation in 23

the surface averaged Cp values is ± 0.12 of the mean result. However, the results present some outliers, as 24

indicated in Figure 1. The variation in the results stresses the importance of quality assurance procedures 25

in wind tunnel experiments. 26

Figure 1 shows that the agreement is better for the windward surface, while the leeward surface and 27

the roof show larger differences. Some possible explanations for the variation in the results were 28

mentioned by the authors, such as ―statistical variability of the data themselves as well as those 29

(8)

introduced by the measurement equipment; physical variability of the flow due to different simulation 1

methods, in particular regarding the structure of the simulated turbulence; different judgement on the time 2

and geometric scales imposed by a given wind-tunnel flow; imperfections of the model, pressure tapping 3

and tubing; imperfections of the software used for the data analysis; and finally, human error, lack of 4

accuracy and ability must not be forgotten.‖ [36]. 5

A comparisons between wind tunnel and full scale results for the same isolated cube case was recently 6

published [37]. It shows an underestimation by wind tunnel experiments at the leeward surfaces. New 7

wind tunnel experiments were conducted in the same study [37], paying special attention to the high-8

frequency part of the turbulence spectra. The results are shown in Figure 2. The new wind tunnel 9

experiment shows much better agreement, even though the full turbulence spectra, and consequently the 10

turbulence intensity profiles, were not reproduced in the wind tunnel. 11

Similarly to the full scale experiments, uncertainty in wind tunnel measurements is seldom published 12

in accordance to the guidelines of ISO [34]. 13

Based on the short sample of studies provided in the section, it can be concluded that wind tunnel 14

experiments present specific challenges. The quality of wind tunnel results is directly affected by the 15

history of calibration in the wind tunnel, quality assurance procedures, and the know-how of the personal 16

involved in the test setup and execution. 17

3.3 CFD 18

CFD has been used to study flow around buildings for more than 30 years [38], while works focused 19

on wind pressure on the building facade become more common about 20 years ago [39-42]. Those studies 20

were clearly exploratory, with no direct application in the building industry. 21

In the next years, CFD use increasingly grew due to the improvements in computer performance, price 22

reduction, and the availability of commercial CFD software. Stathopoulos provided a clear picture of 23

CFD ―past achievements and future challenges‖ in a paper from 1997, which is still equally valid today in 24

many respects [43]. The paper expresses concern about the misuse of CFD for problems that cannot be 25

approached using this technique, which is still the case today considering the lack of validation in several 26

CFD applications. The review indicates some areas for improvement in the future. The list is reproduced 27

below [43]: 28

(9)

―(a) Numerical accuracy by using higher-order approximations coupled with grid independence 1

checks; 2

(b) Boundary conditions, which depend on the specific problem under consideration so that they 3

require good physical insight and high level of expertise; and 4

(c) Refined turbulence models although ad hoc turbulence model modifications are unlikely to 5

perform well beyond the specific flow conditions for which they have been made.‖ [43] 6

It can be said that the advances in these aspects were small compared to the increase in CFD use by 7

practitioners and researchers. Numerical accuracy and grid independence are not reported in many 8

applications, and are not part of the editorial policy in many journals. Concerning the boundary 9

conditions, the definition of wind profiles is often still simplistic. The turbulence profiles are included, 10

but other features in the turbulence structure are neglected in Reynolds-averaged Navier Stokes (RANS) 11

models. In transient simulations, such as those using Large Eddy Simulation (LES), the definition of the 12

boundaries is even more complex. Other features related to the wind profile, such as the profile vertical 13

displacement (zd) are rarely taken into account. The proper use of wall functions for the solid boundary in 14

the domain floor also proved to be a source of concern [44;46]. 15

Finally, there is still no consensus on turbulence modelling for simulation of pressures on facades. It is 16

accepted that LES can provide good results for Cp and for related natural ventilation problems, while 17

RANS simulations present less accurate results [47-49]. Important contributions to the accuracy and 18

reliability in CFD modelling for other applications, such as pedestrian wind comfort and pollutant 19

dispersion, have become available in recent years [50;51]. 20

Despite the vast increase in application of CFD to study the wind flow around buildings, it is not a 21

common practice to use it as source of custom Cp data for BES simulation. The main reasons are the 22

required level of expertise and the high computational cost of these simulations, when compared to the 23

BES simulation itself. Part of those limitations can be overcome in the future by the integration of pre and 24

post processing between BES and CFD, together with clear guidelines for such simulations. IES<VE> is 25

the only BES tool that at present includes a prototype of this integration, but in the present state it is not 26

yet useful, due to the limited range of options for the CFD simulation, limited grid options and lack of 27

integration in the post processing stage. However, it does indicate a possible direction to improve the use 28

of CFD as source of Cp data. 29

(10)

4 Secondary sources: general aspects 1

When primary sources data are not available, secondary sources provide low cost data for infiltration 2

and ventilation studies. Table 2 presents a list of secondary sources implemented in the BES and AFN 3

programs that were mentioned in Table 1. 4

Concerning building height, data for low-rise buildings is more often implemented than data for their 5

high-rise counterparts. Concerning the data source, data provided by AIVC is included in 7 out of the 10 6

programs analysed in this paper. 7

Despite the large amount of wind tunnel data published, only two databases are used in the programs: 8

the so called AIVC database [1;15] and the ASHRAE Handbook - Fundamentals [52]. There seems to be 9

a clear choice for data from ―safe sources‖, supported by well known institutions such as AIVC and 10

ASHRAE. 11

Analytical tools are less frequently included than databases, at least in BES programs. Only two BES 12

programs provide full implementations of the analytical models ―CpCalc+‖ or ―Swami & Chandra‖, 13

while other tools require using third-party analytical tools. The equations proposed by Swami & Chandra 14

[53] are listed in the analytical tools, but they are in fact much simpler then CpCalc+ or Cp Generator. 15

Each of the of secondary data sources in table 2 is discussed in the next sections. 16

5 Secondary sources: Databases 17

Cp databases are compilations of Cp data from one or more sources, where the data is classified 18

according to some parameters, such as building shape and orientation to the incident wind. Cp databases 19

are widely available, particularly for the calculation of wind loads on structures. Wind load standards 20

provide Cp values for unsheltered buildings with simple geometries, to be used when custom wind tunnel 21

experiments are not available. The same approach is used in Cp databases available in the ventilation and 22

infiltration literature, e.g. the AIVC database [1] or the data in the ASHRAE Handbook [52], which are 23

described below. 24

5.1 AIVC 25

The Air Infiltration and Ventilation Centre (AIVC) has been an international reference in the subject 26

since its inauguration in 1979. It is an annex running under the Energy Conservation in Buildings and 27

Community Systems (ECBCS) of the International Energy Agency (IEA). In 1986, after a workshop 28

(11)

about wind pressure coefficients promoted by the AIVC [54], it published a compilation of Cp data as part 1

of a guide [1]. This publication presents tables with data for low-rise buildings, and figures with vertical 2

profiles for high-rise buildings. The tables were compiled based on several studies, while the profiles 3

were reproduced from the original publication [55]. 4

The data for low-rise buildings (up to 3 storeys) is based on the compilation of wind tunnel data 5

published in the workshop [54], and seven other bibliographical references, e.g. [55;56], but the method 6

to compile the database is not mentioned. 7

The Cp database for low-rise buildings consists of tables with surface averaged data, for rectangular 8

floor plans and for 3 shielding levels: exposed, semi-sheltered (obstacles with half of the building height), 9

and sheltered (obstacles with the same height as the building). The data are provided for wind direction 10

sectors of 45°, for a square floor plan building, and for the long and short walls of a rectangular (1:2) 11

floor plan building. The exact building height is not mentioned. For the facades, only the averaged value 12

over the whole surface is provided. For the roof of the low-rise buildings, three types of averaged data are 13

provided: a surface-averaged value, a value for the ―rear‖ and a value for the ―front‖ part of the roof. For 14

each one of them, data is provided according to different roof pitch angle: lower than 10°, between 11° 15

and 30° and higher than 30°. 16

Some details about the data for low-rise buildings are not included in the publication. For the sheltered 17

cases, the spacing between the building and the surroundings obstacles is not mentioned. No information 18

is provided about the wind profile used in the wind tunnel tests. 19

The publication contains several warnings regarding the reliability of the data for low-rise buildings. 20

The first page describes that ―The intention of these data sets is to provide the user with an indication of 21

the range of pressure coefficient values which might be anticipated for various building orientations and 22

for various degrees of shielding.‖ All the other pages contain an explicit warning: ―Caution: Approximate 23

data only. No responsibility can be accepted for the use of data presented in this publication.‖ 24

Despite the modest purpose of the low-rise buildings Cp database, it has been extensively reproduced 25

by AIVC[15;57;58] as well as by other publications about building performance [3;24;59]. The database 26

is currently in use for the scientific community, e.g. [60]. The confidence that is often expressed in this 27

database seems to exceed the intention of the original publication, and in some cases the data are even 28

used to calibrate coefficients in analytical models [61]. 29

(12)

Concerning the AIVC data for high-rise buildings, no effort was made to compile tables based on 1

several wind tunnel tests, and only the data from one source were reproduced in the AIVC publications 2

[1;57]. The data are presented as vertical Cp profiles for two wind directions, 0° and 45° in relation to the 3

normal of the longer face of the model. This angular discretization is commonly used for squared floor 4

plan buildings, but in this case the floor plan has an aspect ratio of 1.5:1 [57], which might lead to 5

misinterpretation by the user. The model used in the wind tunnel tests, at a scale of 1:400, has a height of 6

0.23 m, which represents 92 m in full scale. Data is presented for 4 different shielding levels. The 7

comment provided to contextualize these data mentions that it is ―showing the vertical dependency of 8

pressure coefficients for tall buildings‖, which seems to indicate that the illustrative aspect was a priority, 9

rather than the informative one. 10

The use of vertical profiles for the high-rise building data might lead to misunderstandings, because in 11

some cases the ―vertical dependency‖ is not the main aspect in the Cp distribution over the surface. Figure 12

3 presents an example of vertical Cp profile [1] and the Cp distribution over the same surface for the same 13

experiment [57]. The figure represents a windward facade with a wind attack angle of 45°, and the profile 14

is based on the average of the three values at the same level. The surface distribution shows a clear 15

vertical dependency of Cp, but it also shows a horizontal dependency which is as pronounced as the 16

vertical one for this specific surface and wind attack angle. This dependency is omitted in the vertical 17

profiles, like the one in Figure 3. 18

More recent AIVC publications do not include reproduction of the profiles or surface distributions for 19

rise buildings, and present only the data for low-rise buildings [15;58]. The reproduction of the high-20

rise building data by others is also less common, e.g. [62], and it is used merely to exemplify the complex 21

distribution of Cp over the surface. 22

5.2 ASHRAE 23

The ASHRAE handbook [52] is not a ventilation oriented document like the AIVC publications, so it 24

only presents condensed information about Cp in the chapter dedicated to the airflow around buildings. 25

Different from the AIVC low-rise building database, the ASHRAE handbook only reproduces data from 26

primary sources, rather than compile several data in a single database. 27

(13)

The publication provides data for low and high-rise buildings, presenting examples of Cp distribution 1

over the surface as well as surface averaged data. The building geometries are simple parallelepipeds, 2

with different floor plan aspect ratios and pitch roofs are included with different angular discretizations. 3

An important difference between the AIVC and the ASHRAE handbook data is the attention given to 4

building obstruction effects: ASHRAE does not present data for sheltered buildings, although it provides 5

correction values for the reference wind speed based on sheltering factors. As in the AIVC database, there 6

is no information about the wind profiles used in the experiments. 7

6 Secondary sources: Analytical models 8

The analytical models consist of a set of equations and coefficients to calculate Cp for a specific 9

building configuration [10;11;53;63]. They represent a user-friendly way to access the large amount of 10

empirical data used in the model formulation. Analytical models for Cp prediction were developed based 11

on wind tunnel and full scale experiments. They aim to provide Cp data for a broader range of building 12

configurations, considering obstructions, the effect of different wind profiles and the Cp variation across 13

the facade. None of the models presented here provide the uncertainty in their predictions. For some of 14

them, correlation coefficients are provided, but it is not possible to calculate the prediction uncertainty 15

using only this value. Therefore, it is not possible to assess the quality of their results to predict Cp values 16

for new building configurations. 17

Analytical models were developed using regression techniques to analyse a large amount of Cp data. 18

The result is a function where the Cp value depends on a set of parameters considered in the regression, 19

e.g., facade aspect ratio, position at the facade, building aspect ratio, position and size of the surrounding 20

buildings, wind direction and aerodynamic roughness. The applicability of the derived functions depends 21

on the quality and variety of the Cp experimental data used in the regression, as well as on the parameters 22

considered in the analysis. Regarding the experimental data, several authors [10;53] point to the lack of 23

data for complex shapes such as L-shape or U-shape. Regarding the parameters, two considerations are 24

important. Firstly, the available data guide the parameterization, because the chosen parameter needs to 25

be covered by the range of experiments. So, some parameters cannot be considered because of lack of 26

data. Secondly, there is a trade-off between precision and complexity. More precise equations tend to 27

demand more parameters, but the increment in the precision does not necessarily justify the use of very 28

(14)

complex formulae. The correct choice of which parameters to include in the regression analysis can be 1

made based on sensitivity analysis. 2

In the following sections, the main features of three analytical models are presented. 3

6.1 The model by Swami & Chandra (1988) 4

The model proposed by Swami & Chandra [53] provides one simple equation for low-rise buildings 5

and another for high-rise buildings. The low-rise building equation is presented below, as an example: 6 2 p NC ln[1.248 0.703 sin( / 2) 1.175 sin ( / 2)      7 3 4 0.131 sin ( / 2) 0.769 sin ( / 2)       8 2 5 6 0.07 G sin ( / 2) 0.717 sin ( / 2)]        (2) 9

where NCp is the normalized pressure coefficient, G is the natural logarithm of the floor plan aspect 10

ratio and  11

The equation adopts a normalized pressure coefficient NCp, considering NCp equal to 1 when the wind 12

is orthogonal to the surface. It is therefore necessary to know a priori the Cp value for wind orthogonal to 13

the surface. The model suggests Cp = 0.6 for this case, but it indicates that the value can vary from 0.19 to 14

0.91, depending on the wind profile, building height, roof pitch and floor plan aspect ratio. 15

The model calculates Cp for building with rectangular floor plan, and sheltering effects are not 16

considered in detail. A shielding correction factor is proposed, to be applied directly to the calculated 17

flow rate. 18

The equation for low-rise buildings provides surface averaged Cp. The decision to neglect the 19

variation of Cp over the surface was based on early studies [56;64], which focused on infiltration 20

calculations assuming cracks that are homogeneously distributed over the building facades. This 21

assumption however is not valid for many ventilation calculations, and may also be invalid for some 22

infiltration calculations. The equation has two parameters: wind direction and building floor plan aspect 23

ratio, and has a correlation coefficient of 0.8. This result is good, considering the broad range of data 24

analysed, including data from buildings with different heights and different roof pitch angles, and that fact 25

that these parameters are not used in the analytical model. The equation for high-rise buildings does not 26

provide surface averaged values but includes the position at the facade as an additional parameter. 27

Correlation coefficients are not provided. 28

(15)

6.2 CpCalc+ (1992) 1

CpCalc+ [10] was developed within the COMIS workshop [4] and the European project AIOLOS 2

[24], with the intention to provide results for sheltered buildings, that could not be obtained with the 3

model by Swami & Chandra. Compared to this model [53], further experimental data were used to take 4

into account sheltering effects, and also new parameters were added, such as the power-law exponent of 5

the mean wind velocity profile, the plan area density, the relative building height to the surrounding 6

buildings, the frontal aspect ratio and the position at the facade. Sheltering effects are considered using 7

the ―plan area density‖, where the individual sheltering effects of each building are not taken into 8

account. In order to allow for the higher number of parameters, a parametrical approach was used, based 9

on successive independent corrections for each parameter. This approach creates a much more complex 10

model, based on several tables with coefficients for each correction. The author mentions that the 11

methodology is the main result, rather than the equations themselves. This is due to the lack of consistent 12

experimental data, which is considered the main obstacle for a more comprehensive analysis. 13

6.3 Cp Generator 14

The Cp Generator has been developed in the last 30 years to ―predict the wind pressure coefficients, 15

Cp, on the facades and roofs of block shaped buildings‖ [11]. The tool is a web-based application 16

(http://cpgen.bouw.tno.nl), developed by the Dutch institution TNO. The Cp Generator has a similar 17

approach and similar capabilities as CpCalc+, predicting point values on the facade and also on the roof, 18

dealing with low-rise and high-rise buildings with user-defined dimensions (length-width-height) and 19

taking into account the effects of the surrounding terrain by wind profile corrections. The main 20

improvement is situated in the way in which sheltering is taken into account. The model considers 21

discrete block shaped obstructions instead of the neighbourhood plan area density. Unfortunately, the 22

model was not developed in the English language. A comparison between Cp Generator results and 23

experimental data for that specific low-rise building shows that they ―are closer to reality than simulations 24

carried out with Cp values from the AIVC tables‖ [33]. Nevertheless, the results show large deviations 25

from full scale experimental data for some points and wind directions [33]. 26

7 Comparison of Cp data features

(16)

Table 3 presents the summary of the characteristics described in the previous section. The first 1

conclusion is that none of the databases or analytical methods can handle the effects of the site 2

topography, building facade detailing or inform about the uncertainty of the provided data. Another 3

feature that is similar in most data sources is the size of the wind direction sectors. In most sources, the 4

user can choose the number of the wind direction sectors. Only the AIVC database has a fixed angular 5

discretization in 45° intervals. The variation across the facade and sheltering effects are treated in 6

different ways by each data source presented in Table 1. These aspects are discussed in the subsections 7

below. 8

7.1 Variation across the facade 9

Table 3 shows that the variation across the facade is only considered only by some analytical methods, 10

but most of the BES programs adopt only database values. In order to analyse the importance of the 11

variation across the facade, some comparisons are made here. An arbitrary terrain type, suburban 12

environment, is chosen for this comparison, and the corresponding wind profile parameters for the 13

analytical models are obtained in the program documentation. A power-law wind profile exponent (of 14

0.22 was used for CpCalc+, while an aerodynamic roughness length (z0) of 0.5 m was used for the Cp 15

Generator. For the AIVC database and the model by Swami & Chandra for low-rise buildings, only 16

surface averaged values are used in the comparisons, because those data sources do not provide values for 17

specific points at the facade. Figure 4 presents the data for three points on the facade of a low-rise cubic 18

building (10x10x10 m³). The data from the different sources show a similar pattern, and the range of 19

deviation in the results for the middle point (A) is 0.4, which might be considered high, given the simple 20

building geometry. For the lower left corner (x = 1 m, y = 1 m) the data again show a similar pattern, but 21

the deviations go up to 0.5 for a wind attack angle of 80°. 22

Figure 4 (D) shows the differences in Cp values compared with the surface averaged values of the 23

AIVC. As reported by Ref [65], the use of surface averaged values was one of the main motivations for 24

the development of analytical methods, as ―From experience we know that wall-averaged values of Cp 25

usually do not match the accuracy required for air flow calculation models.‖ 26

7.2 Sheltering effects 27

Table 3 shows the differences in approaches to model the sheltering effect. These approaches can be 28

classified in three categories. The first category considers each surrounding building individually to 29

(17)

provide the combined sheltering effect (marked with ―x‖ in Table 3). The second category considers an 1

averaged effect of the surrounding buildings, with just a global description of the obstacle height and 2

horizontal distribution (marked with ―p‖). The last category adopts simplified correction factors (marked 3

with ―s‖). In order to present the different results obtained using categories one and two, the same 4

building and facade positions of Figure 4 were used, but now this building is considered to be surrounded 5

by similar buildings in an infinite regular array with a horizontal spacing of 10 m. The results are shown 6

in Figure 5. While the AIVC data display only a small variation of Cp values with the angle of attack, the 7

analytical methods provide very different results as a function of this angle. This might seem logical, 8

given the effect of channelling of wind flow through the building group. The range of the results is wide, 9

implying high uncertainty associated with Cp values for sheltered buildings. Figure 5 only considers 10

sheltering by neighbouring buildings, but several low-rise buildings, such as L-shape or U-shape 11

buildings, can provide shelter to themselves [53]. In those cases, the uncertainty might be even higher. 12

Table 4 details the method to calculate the sheltering effect as implemented in each program. The first 13

three lines describe the three data sources used in Figure 5. The first option is to use Cp results from wind 14

tunnel experiments, like in the AIVC database. Another solution is the use of parameters to describe the 15

surrounding area, like the ―neighborhood density‖ in CpCalc+ or the position of a ―discrete obstruction‖ 16

in Cp Generator. Those methods try to take into account the specific impact on the Cp value, instead of 17

adopting a more pronounced simplification like the other methods listed in the table. The simplified 18

correction factor models are described below. 19

IES <VE> adopts the AIVC database, but the data for high-rise sheltered buildings do not comply 20

with the original source. Data analysis revealed that arbitrary correction factors (0.66 and 0.33) are 21

applied to the data for fully exposed buildings to obtain values for semi-exposed and sheltered buildings 22

respectively. The documentation does not mention this procedure. SUNREL adopts a similar method, 23

applying correction factors according to the shielding classification. Tas assumes that Cp for any sheltered 24

facade is equal to the Cp value from the leeward facade of the unobstructed building [19]. The orientation 25

of the sheltered facade relative to the wind direction is not taken into account. The comparison of leeward 26

data in Figure 4 with sheltered data in Figure 5 shows that this approach is not preferable, as it can 27

provide underpressure instead of overpressure. ASHRAE [52] describes a method to take into account the 28

sheltering effect by correcting the reference wind speed. This method is not implemented in any of the 29

(18)

studied programs. All those methods consider a uniform sheltering correction for all points at the building 1

facades, while in reality the extent of sheltering can be different for different parts of the facade. Stand-2

alone AFN tools, which are focused on the ventilation modelling, do not include those simplifications, 3

which can be interpreted as an indication of the less good performance of those approaches. 4

8 Conclusions 5

The present research provided an overview of wind pressure coefficient data in building energy 6

simulation programs. Some points can be highlighted from the findings described in the previous 7

sections. 8

Warnings in the original data sources are not reproduced with the data, which might lead to 9

misunderstandings when using the data. 10

Databases are the most common data source in BES. Analytical tools are rarely found and are poorly 11

integrated with BES programs. CFD is still a promise for the future regarding its integration with BES 12

tools. 13

Data describing the same building present large variations depending on the data source, even for 14

simple configurations like fully exposed cubic buildings. Sheltered buildings and especially points near 15

the facade corners present even higher variations. The same applies to complex building geometries, 16

which are not included in existing secondary databases. 17

The uncertainty associated with the analysed data sources is high. Therefore, the quantification of 18

those uncertainties using empirical data for a broad range of cases is an important topic of future research. 19

This overview may be used to guide future efforts in the development of BES and HAM programs. It 20

will also assist future studies dealing with ventilation simulation, particularly those focused on the impact 21

of Cp data sources in the overall simulation uncertainty 22

Acknowledgements 23

This research is funded by the ―Institute for the Promotion of Innovation by Science and Technology 24

in Flanders‖ (IWT-Vlaanderen) as part of the SBO-project IWT 050154 ―Heat, Air and Moisture 25

Performance Engineering: a whole building approach‖. Their financial contribution is gratefully 26

acknowledged. 27

(19)

References 1

[1] M.W. Liddament, Air infiltration calculation techniques - an applications guide, AIVC, Bracknell, 2

1986. 3

[2] C. Hagentoft, Heat, air and moisture transfer in insulated envelope parts: Performance and 4

Practice, International Energy Agency, Annex 24. Final Report, vol. 1. Acco, Leuven, 1996. 5

[3] J. A. Clarke, Energy simulation in building design, Butterworth-Heinemann, Oxford, 2001. 6

[4] H.E. Feustel, COMIS — an international multizone air-flow and contaminant transport model, 7

Energy and Buildings 30 (1999) pp. 3-18. 8

[5] N. Sahal, M. Lacasse, Water entry function of a hardboard siding-clad wood stud wall, Building 9

and Environment 40 (2005) pp. 1479-1491. 10

[6] H. Hens, A. Janssens, W. Depraetere, J. Carmeliet and J. Lecompte, Brick Cavity Walls: A 11

Performance Analysis Based on Measurements and Simulations, Journal of Building Physics 31 (2007) 12

pp. 95-124. 13

[7] S. de Wit, G. Augenbroe, Uncertainty analysis of building design evaluations, Proceedings of the 14

7th Internationl IBPSA Conference, Rio de Janeiro, 2001. 15

[8] S. de Wit, Uncertainty in predictions of thermal comfort in building, PhD Thesis, Delft University 16

of Technology, 2001. 17

[9] J.L.M. Hensen, On the thermal interaction of building structure and heating and ventilating 18

system, PhD Thesis, Eindhoven University of Technology, 1991. 19

[10] M. Grosso, Wind pressure distribution around building: a parametrical mode, Energy and 20

Building 18 (1992) pp. 101-131. 21

[11] B. Knoll, J.C. Phaff, W.F. de Gids, Pressure simulation program, Proceedings of the Conference 22

on Implementing the Results of Ventilation Research, AIVC, 1995. 23

[12] P. Tuomaala, Implementation and evaluation of air flow and heat transfer routines for building 24

simulation tool, PhD Thesis, Helsinki University of Technology, 2002. 25

[13] D. Crawley, J.W. Hand, M. Kummert, B. Griffith, Contrasting the capabilities of building energy 26

performance simulation programs, Building and Environment 43 (2008) pp. 661-673. 27

[14] A. Haas, A. Weber, V. Dorer, W. Keilholz, R. Pelletret, COMIS v3.1 simulation environment for 28

multizone air flow and pollutant transport modeling, Energy and Buildings 34 (2002) pp. 873-882. 29

(20)

[15] M.L. Orme, Technical Note 51 - Applicable models for air infiltration and ventilation 1

calculations, AIVC, 1999. 2

[16] ESP-r source code, http://www.esru.strath.ac.uk/Programs/ESP-r_central.htm, accessed in 3

17.03.2008. 4

[17] EnergyPlus, EnergyPlus Engineering Reference - The Reference to EnergyPlus Calculations, 5

2007. 6

[18] Integrated Environmental Solutions, MacroFlo Calculation Methods - <Virtual Environment> 7

5.6, n.d. 8

[19] EDSL, TAS Theory, EDSL, n.d. 9

[20] K.T. Andersen, Natural ventilation model in BSim, Danish Building Research Institute, 2002. 10

[21] K.B. Wittchen, K. Johnsen, K. Grau, BSim User's Guide, Danish Building Research Institute, 11

2004. 12

[22] M. Deru, R. Judkoff, P. Torcellini, SUNREL Technical Reference Manual, NREL, Golden, 2002. 13

[23] G.N. Walton, W.S. Dols, CONTAM 2.4b User Guide and Program Documentation, NIST, 14

Gaithersburg, 2006. 15

[24] F. Allard (ed.), Natural ventilation in buildings: a design handbook, James x James, London, 16

1998. 17

[25] M. Jensen, N. Franck, Model-scale tests in turbulent wind - Part II, Copenhagen, The Danish 18

technical press, 1965. 19

[26] P.J. Richards, R.P. Hoxey, L.J. Short, Wind pressures on a 6m cube. J. Wind Eng. Ind. Aerodyn. 20

89 (2001) pp. 1553–1564. 21

[27] G. M. Richardson, A. P. Robertson, R. P. Hoxey, D. Surry, Full-scale and Model Investigations 22

of Pressures on an Industrial/Agricultural Building. J. Wind Eng. Ind. Aerodyn., 36 (1990) pp. 1053-23

1062. 24

[28] M.L. Levitan, J.D. Holmes, K.C. Mehta and W.P. Vann, Field measurements of pressures on the 25

Texas Tech building. J. Wind Eng. Ind. Aerodyn. 38 (1991) pp. 227–234. 26

[29] T. A. Reinhold, Wind tunnel modeling for civil engineering applications. Cambridge University 27

Press, Cambridge, 1982. 28

(21)

[30] D.W. Etheridge, Unsteady flow effects due to fluctuating wind pressures in natural ventilation 1

design - mean flow rates, Building and Environment 35 (2000), pp. 111-133. 2

[31] D. Cóstola, D.W. Etheridge, Unsteady natural ventilation at model scale—Flow reversal and 3

discharge coefficients of a short stack and an orifice, Building and Environment 43 (2008) pp. 1491-1506. 4

[32] R. Stull, An introduction to boundary layer meteorology, Kluwer Academic Publishers, 5

Dordrecht, 1988. 6

[33] N. Heijmans, P. Wouters, IEA Technical report: Impact of the uncertainties on wind pressures on 7

the prediction of thermal comfort performances - HybVent Annex35, n.d. 8

[34] ISO. Guide to the expression of uncertainty in measurement (GUM), ISO, Switzerland, 1995. 9

[35] J.O.V. Irminger, C.H.R. Nøkkentved, Wind-pressure on buildings, Copenhagen, Danmarks 10

Naturvidenskabelige Samfund, 1936. 11

[36] N. Holscher, H.J. Niemann, Towards quality assurance for wind tunnel tests: A comparative 12

testing program of the Windtechnologische Gesellschaft. J. Wind Eng. Ind. Aerodyn. 74 (1998), pp. 599-13

608. 14

[37] P.J. Richards, R.P. Hoxey, B.D. Connell, D.P. Lander, Wind-tunnel modelling of the Silsoe 15

Cube, J. Wind Eng. Ind. Aerodyn. 95 (2007) pp. 1384–1399. 16

[38] C. W. Hirt, J. L. Cook, Calculating three-dimensional flows around structures and over rough 17

terrain, Journal of Computational Physics 10 (1972) pp. 324-340. 18

[39] E. H. Mathews, Prediction of the wind-generated pressure distribution around buildings, J. Wind 19

Eng. Ind. Aerodyn. 25 (1987) pp. 219-228. 20

[40] S. Murakami, A. Mochida, K. Hibi, Three-dimensional numerical simulation of air flow around a 21

cubic model by means of large eddy simulation, J. Wind Eng. Ind. Aerodyn. 25 (1987) pp. 291-305. 22

[41] S. Murakami, A. Mochida, 3-D numerical simulation of airflow around a cubic model by means 23

of the k-ε model, J. Wind Eng. Ind. Aerodyn. 31 (1988) pp. 283-303. 24

[42] S. Murakami, Computational wind engineering, J. Wind Eng. Ind. Aerodyn. 36 (1990) pp. 517-25

538. 26

[43] T. Stathopoulos, Computational wind engineering: Past term achievements and future challenges, 27

J. Wind Eng. Ind. Aerodyn. 67 (1997) pp. 509-532. 28

(22)

[44] B. Blocken, T. Stathopoulos, J. Carmeliet, CFD simulation of the atmospheric boundary layer: 1

wall function problems. Atmospheric Environment 41 (2007), pp. 238-252. 2

[45] B. Blocken, J. Carmeliet, T. Stathopoulos, CFD evaluation of wind speed conditions in passages 3

between parallel buildings—effect of wall-function roughness modifications for the atmospheric 4

boundary layer flow, J. Wind Eng. Ind. Aerodyn. 95 (2007) pp. 941-962. 5

[46] D.M. Hargreaves, N.G. Wright, On the use of the k–ε model in commercial CFD software to 6

model the neutral atmospheric boundary layer , J. Wind Eng. Ind. Aerodyn. 95 (2007), pp. 355-369. 7

[47] Y. Tominaga, A. Mochida, S. Murakami, S. Sawaki, Comparison of various revised k–ε models 8

and LES applied to flow around a high-rise building model with 1:1:2 shape placed within the surface 9

boundary layer, J. Wind Eng. Ind. Aerodyn. 96 (2008), pp. 389-411. 10

[48] T. Yang, N.G. Wright, D. Etheridge, A. Quinn, A comparison of CFD and full-scale 11

measurements for analysis of natural ventilation, The International Journal of Ventilation 4 (2006) pp. 12

337-348. 13

[49] T. Nozu, T. Tamura, Y. Okuda, S. Sanada, LES of the flow and building wall pressures in the 14

center of Tokyo. J. Wind Eng. Ind. Aerodyn. (2008) doi:10.1016/j.jweia.2008.02.028 15

[50] J. Franke, A. Hellsten, H. Schlünzen, B. Carissimo, Cost Action 732 - Best practice guideline for 16

the CFD simulation of flows in the urban environment, Brussels, COST 2007. 17

[51] Y. Tominaga, A. Mochida, R. Yoshie, H. Kataoka, T. Nozu, M. Yoshikawa, T. Shirasawa, AIJ 18

guidelines for practical applications of CFD to pedestrian wind environment around buildings, J. Wind 19

Eng. Ind. Aerodyn. 96 (2008) pp. 1749–1761. 20

[52] ASHRAE, ASHRAE Handbook - Fundamentals, ASHRAE, Atlanta, 2001. 21

[53] M.V. Swami, S. Chandra, Correlations for pressure distribution on buildings and calculation of 22

natural-ventilation airflow, ASHRAE Transactions 94 (1988) pp. 243–266. 23

[54] AIVC, Wind Pressure Workshop Proceedings AIC-TN-13.1-84, AIVC, Brussels, 1984. 24

[55] J.J. Bowen, A wind tunnel investigation using simple building models to obtain mean surface 25

wind pressure coefficients for air infiltration estimates - NRC Report LTR-LA-209, National Research 26

Council, Canada, 1976. 27

[56] B.G. Wiren, Effects of surrounding buildings on wind pressure distributions and ventilation 28

losses for single family houses - M85:19, National Swedish Institute for Building Research, Gavle, 1985. 29

(23)

[57] M.L. Orme, M.W. Liddament, A. Wilson, Technical Note 44 - Numerical data for air infiltration 1

and natural ventilation calculations, AIVC, 1998. 2

[58] M.L. Orme, N. Leksmono, AIVC Guide 5, AIVC, 2002. 3

[59] M. Santamouris, Passive cooling of building, James x James, London, 1996. 4

[60] O.S. Asfour, M.B. Gadia, A comparison between CFD and network models for predicting wind-5

driven ventilation in buildings, Building and Environment 42 (2007) pp. 4079-4085. 6

[61] G. van Moeseke, E. Gratia, S. Reiter, A. Herde, Wind pressure distribution influence on natural 7

ventilation for different incidences and environment densities, Energy and Buildings 37 (2007) pp. 878-8

889. 9

[62] M. Santamouris, P. Wouters, Building ventilation – the state of the art, Earthscan, London, 2006. 10

[63] A.S. Eldin, A parametric model for predicting wind-induced pressures on low-rise vertical 11

surfaces in shielded environments, Solar Energy 81 (2007), pp. 52-61. 12

[64] M.V. Swami, S. Chandra, Procedures for calculating natural ventilation airflow rates in buildings 13

- Final Report FSEC-CR-163-86, Florida Solar Energy Center, Cape Canaveral, 1987. 14

[65] H.E. Feustel B.V. Smith, V. Dorer, A. Haas, A. Weber, COMIS 3.2 - User Guide, Empa, 15

Dubendorf, 2005. 16

17 18

(24)

Table 1 . BES and AFN program 1

Program Type Version Documentation ESP-r BES 11.3 [3;9;16] EnergyPlus 2.0.0.025 [17] IES <VE> 5.6 [18] Tas 9.0.5 [19] BSim 4.6.7.12 [20;21] IDA ICE 3.0 (15) - SUNREL - [22] CONTAMW AFN 2.4b [23] AIOLOS 1.0 [24] COMIS 3.2 [4;14] 2

(25)

Table 2 . Cp database and analytical models implemented in BES and AFN software 1 E SP -r E n er g y Plu s IE S < VE > T as B Sim IDA IC E SUNR E L C ONT AM W AI OL OS C OM IS Database AIVC Low-rise x x p p ? ? High-rise p p ? ? ASHRAE Low-rise High-rise x ? Analytical model Swami & Chandra Low-rise x p p2 High-rise CpCalc+ x x x1 Cp Generator x1 x1 x – Implemented p – Partially implemented

? – Not clearly mentioned in the documentation Notes:

1. Not implemented in the software. It is indicated as Cp source in the software documentation.

2. Table with results based on the analytical equation. 2

(26)

Table 3 . Database and analytical model features 1 Prim ar y s o u rce s AI VC ASHR AE Swam i & C h an d ra C p Gen er ato r L o w -r is e Hig h -r is e L o w -r is e Hig h -r is e L o w -r is e Hig h -r is e C p C alc+ Topographic effects x

Effect of surrounding terrain

(smooth, rural, suburban, urban) x x x Effect of immediate surroundings

(local sheltering by buildings, etc) x p p s s s s p x Building configuration (building

geometry and facade detailing) x

Size of direction sectors (degrees) c 45 90 c c c c c c Variation across the facade x s x x x

Uncertainty x

x – Fully able to model p – Partially able to model s – Simplified model c – Custom

(27)

Table 4 . Methods to model sheltering effects 1 E SP -r E n er g y Plu s IE S < VE > T as B Sim IDA IC E SUNR E L C ONT AM W AI OL OS C OM IS Cp x x x x x x x Neighborhood density x x x Discrete obstruction x x Correction factor on Cp x

Copy data from the leeward facade x

Correction factor on flow rate x

Correction factor on wind speed 2

(28)

1

Figure 1 . Comparison of different wind tunnel experiments, after Ref. [36]. 2

3 4 5

6

Figure 2 . Comparison of different wind tunnel experiments with full scale results, after Ref. [36;37]. 7

8 9 10

(29)

1

Figure 3 . Example of vertical profile of Cp values for a high-rise building surface [1] (left), and the Cp 2

distribution over the same surface [57] (right). 3

4 5 6

(30)

1

Figure 4 . Cp for an unsheltered low-rise cubic building as function of wind angle of attack (): (A) 2

middle of the facade, (B) lower left corner, (C) upper right corner, (D) range of data for the points A, B 3 and C. 4 5 6 7 8 9 10 11 12 13 14

(31)

1 2

3

Figure 5 . Sheltered building: Cp from different sources, and different points in the facade. 4

Referenties

GERELATEERDE DOCUMENTEN

Daarom werd een prospectie met ingreep in de bodem aanbevolen, zodat een inschatting kan gemaakt worden van eventueel op het terrein aanwezige archeologische waarden,

Prediction of fatigue crack growth for welded joints using stress intensity factors determined by FEM calculations Citation for published version (APA):..

The aim of the study was to determine second-year physiotherapy students’ perceptions about the learning opportunities provided in the introductory clinical module and to

Initially there is a fast deactivation of the catalyst, and after a long time a constant rate of reaction is obtained. By fitting the curves presented in

Indien mogelijk kunt u uw afspraak voor het MRI onderzoek zodanig proberen te plannen dat deze op een gunstig tijdstip van vervanging van de sensor valt... Voorbereiding thuis

For the purpose of this study patient data were included based on the following criteria: (1) consecutive adults who underwent a full presurgical evaluation for refractory

Bovendien staat CM loodrecht op CN (binnen- en buitenbissectrice van dezelfde hoek). De constructie kan nu als volgt verlopen... 1) Construeer de (congruente)

side initiatives Growth in energy demand exceeding electricity supply European Union Emission Trading Scheme (EU ETS) Clean Development Mechanism (CDM) Eskom – South