Anda di halaman 1dari 100

Numerical analysis of thermal bridges under dynamic

boundary conditions

Stijn Van Nieuwenhove

Supervisors: Prof. dr. ir. arch. Arnold Janssens, dr. ir. arch. Nathan Van Den Bossche
Dr. Hua Ge

Master`s dissertation submitted in order to obtain the academic degree of


Master of Science: Architecture

Department of Architecture and Urban planning


Chairman: Prof. dr. Pieter Uyttenhove
Faculty of Engineering and Architecture
Academic year 2013-2014
Permission for use of content

The author gives permission to make this master dissertation available for consultation and to copy parts
of this master thesis for personal use. In the case of any other use, the limitations of the copyright have to
be respected, in particular with regard to the obligation to state expressly the source when quoting
results from this work.

Montréal, 2 June 2014


The author

Stijn Van Nieuwenhove

Toelating tot bruikleen

De auteur geeft de toelating deze masterproef voor consultatie beschikbaar te stellen en delen van de
masterproef te kopiëren voor persoonlijk gebruik. Elk ander gebruik valt onder de beperkingen van het
auteursrecht, in het bijzonder met betrekking tot de verplichting de bron uitdrukkelijk te vermelden bij
het aanhalen van resultaten uit deze masterproef.

Montréal, 2 juni 2014


De auteur

Stijn Van Nieuwenhove

iii
iv
Numerical analysis of thermal bridges under dynamic
boundary conditions

Stijn Van Nieuwenhove

Supervisors: Prof. dr. ir. arch. Arnold Janssens, dr. ir. arch. Nathan Van Den Bossche
Dr. Hua Ge

Master`s dissertation submitted in order to obtain the academic degree of


Master of Science: Architecture

Department of Architecture and Urban planning


Chairman: Prof. dr. Pieter Uyttenhove
Faculty of Engineering and Architecture
Academic year 2013-2014

v
Acknowledgements

First of all, I would like to thank Hua Ge and Nathan Van Den Bossche in particular, for supervising me
during the experimental journey of this thesis and would like to express my gratitude to them and Arnold
Janssens for giving me the opportunity to perform my thesis at Concordia University in Montreal. I want to
thank Kim Carbonez for her advice on the Voltra software.

I also want to thank the people in the building envelope office; Colin, Ramy, Clement, Angel, Vincent and
Stephanie for their kindness and helpfulness during my staying at Concordia. Because they enlightened me
about their own experiments and involved me in other amazing side-activities, I’ve learned a lot.

I want to express my gratitude to all my friends and family for their support and distraction they provided
me during my studies. Especially my parents, for given me this opportunity and their unconditional support
and believe in me.

Last but not least, I would like to thank my girlfriend for her endless love and support she has given me
through my study.

Stijn Van Nieuwenhove, 2 June 2014

vi
Numerical analysis of thermal bridges under dynamic boundary conditions

By Stijn Van Nieuwenhove

Supervisors: prof. dr. ir. Arnold Janssens, dr. ir Nathan Van Den Bossche, dr. ir Hua Ge

Master thesis submitted to obtain the academic degree in Master of Science: Architecture

Department of architecture and urban planning


Chairman: Prof. dr. Pieter Uyttenhove
Faculty of engineering and architecture
University of Ghent
Academic 2013-2014

Abstract

In Northern America there is al lot of research going on about thermal bridges and building envelope
interfaces. These are, in contrast to Belgium, not included in the standard calculations for heat losses. The
considerable thermal bridges such as balcony slabs are present in many existing and newly-constructed
multi-unit residential buildings (MURBs) in Canada. In this thesis, building envelope interfaces are
simulated in transient boundary conditions and compared to a steady state method that calculates the
total heat loss of a building envelope interface. By using the linear thermal transmittance factor, 2D heat
losses are included in the calculations. The comparison between transient and steady state calculations
gives an idea whether or not the linear thermal transmittance calculated under steady state conditions is
a good approximation.

A valuable tool to detect thermal bridges and the related heat losses in building envelopes is the use of
infrared thermography. The influence of different boundary conditions is investigated by using numerical
simulations performed by the thermal computer software Voltra. For each time interval of the transient
simulations, the surface temperature is reported and the temperature factor is calculated. These dynamic
temperature factors are compared to those calculated in steady state conditions for a specific location.

vii
The aim of this analysis is to evaluate whether or not thermography can be used to detect thermal
bridges (qualitative assessment), or even to get a rough estimate of the temperature factor (quantitative
assessment). Since the temperature factor calculated under static boundary conditions is often used to
assess the risk for mould growth, a reliable estimate using thermography could be a great asset for
planning building retrofits. Further on, guidelines are compiled to perform an infrared thermography
survey in more reliable conditions.

Keywords:
Linear thermal transmittance factor, heat transfer in transient conditions, thermal bridges, Infrared
thermography, Voltra simulations.

viii
Numerieke analyse van koudebruggen onder dynamische randvoorwaarden

Door Stijn Van Nieuwenhove

Promotors: prof. dr. ir. Arnold Janssens, dr. ir Nathan Van Den Bossche, dr. ir Hua Ge

Master thesis ingediend tot het behalen van de academische grad van Master in de
ingenieurswetenschappen: architectuur

Vakgroep Architecture en Stedenbouw


Voorzitter: Prof. dr. Pieter Uyttenhove
Faculteit Ingenieurswetenschappen en Architectuur
Academiejaar 2013-2014

Samenvatting

In Noord-Amerika worden thermische bruggen en bouwknopen intensief onderzocht. In tegenstelling tot


België, zijn deze niet standaard ingecalculeerd voor hitteverliezen. Veel bestaande en nieuwe multi-unit
residentiële gebouwen in Canada bezitten veel balkons met de bijhorende koude bruggen. In deze thesis
zijn bouwknopen gesimuleerd in dynamische omgevingsfactoren en vergeleken met de statische
methoden die de totale warmte verliezen van bouwknopen berekenen. Gebruik makend van de lineaire
transmissie factor, 2D hitte verliezen zijn in rekening opgenomen. De vergelijking tussen de dynamische
en statische berekeningen geven een idee of de lineaire warmte transmissie berekend via de statsiche
methode een goede benadering is.

Infrarood thermografie is een waardevol hulpmiddel om deze koude bruggen en de gerelateerde warmte
verliezen in bouwknopen the detecteren. De invloed van de verscheidene omgevingsfactoren is
onderzocht met numerieke simulaties met het thermische computerprogramma Voltra. Voor elk
tijdsinterval is de oppervlaktetemperatuur gerapporteerd en de temperatuurfactor berekend. Deze
dynamische temperatuursfactoren werden vergeleken met deze berekend in statische condities op een
specifieke locatie. Het doel van deze analyse is bepalen of infrarood thermografie al dan niet gebruikt kan

ix
worden om koude bruggen te detecteren (kwalitatief) of zelfs een ruwe schatting te geven van de
temperatuursfactor (kwantitatief). Aangezien de temperatuurfactor berekend onder statische
omgevingsfactoren vaak gebruikt wordt om het risico voor schimmelgroei in te schatten, kan dit een
groot voordeel zijn voor het bepalen ervan in gebouwen. Richtlijnen zijn samengesteld die aanbevelingen
geven om een thermografisch onderzoek uit te voeren in de meest gunstige condities.

Trefwoorden
Lineaire transmissie coëfficiënt, warmte overdracht in dynamische omstandigheden, koudebruggen,
infraroodthermografie, Voltra simulaties

x
Table of Contents

PERMISSION FOR USE OF CONTENT III


TOELATING TOT BRUIKLEEN III
AKNOWLEDGEMENTS VI
SUMMARY VII
SAMENVATTING IX
TABLE OF CONTENTS XI
CHAPTER 1 –INTRODUCTION 1
1.1 CONTEXT 1
1.2 OBJECTIVES 3
CHAPTER 2 – HEAT TRANSFER IN BUILDING COMPONENTS 5
2.1 Introduction 5
2.2 Thermal performance in steady state conditions 6
2.2.1 Heat transfer calculation 6
2.2.2 Solar radiation 11
2.2.3 Surface heat transfer coefficient 12
2.3 Transient heat transfer in building components 14
2.3.1 Dynamic boundary conditions 15
2.3.2 The effect of insulation location on the dynamic heat transfer 17
2.3.3 Different methods for dynamical building simulations 18
2.4 Design guidelines for building envelope interfaces 20
2.4.1 Influence of mould and surface condensation on building components 20
2.4.2 Thermal performance of building components 22
CHAPTER 3 – SIMULATION MODEL AND NUMERICAL ANALYSIS 25
3.1 Introduction 25
3.2 Calculation of heat transfer in Voltra 25
3.2.1 Input 26
3.2.2 Output 29
3.3 Building junctions 30
3.3.1 Wall/roof junction 31
3.3.2 Balcony connection 33
3.3.3 Wall/floor junction 35
3.4 Model assumptions 36
3.4.1 Boundary conditions 36
3.4.2 Calculation method 38
3.5 Comparison of the steady state and the dynamic heat loss calculation 42
3.6 Conclusions 43
CHAPTER 4 – INFRARED THERMOGRAPHY 45
4.1 Introduction 45

xi
4.2 Use of infrared thermography 46
4.2.1 accuracy of infrared pictures 47
4.3 Influence factors on infrared thermography 48
4.3.1 General 48
4.3.2 Environmental factors 49
4.4 Conclusions 51
CHAPTER 5 – VALIDATION OF GUIDELINES TO DETECT DIFFERENT BUILDING ENVELOPE 53
INTERFACES
5.1 Introduction 53
5.2 Model assumptions 54
5.2.1 General 54
5.2.2 Boundary conditions 55
5.2.3 Calculation method 55
5.3 Wall/roof junction 57
5.3.1 General 57
5.3.2 Influence of the wind speed on the internal temperature factor 59
5.3.3 Influence of the temperature difference between indoor and outdoor on the
internal temperature factor 61
5.3.4 Influence of the solar radiation on the internal temperature factor 63
5.3.5 Influence of al guidelines on the internal surface temperature 65
5.3.6 Guidelines applied on all wall/roof junctions 66
5.4 Balcony connections 68
5.4.1 Measurements on the internal surface 68
5.4.2 Measurements on the external surface 70
5.4.3 Influence of the wind speed on the external temperature factor 72
5.4.4 Influence of the temperature difference between indoor and outdoor 72
environment on the external temperature factor
5.4.5 Influence of the solar radiation on the external temperature factor 74
5.4.6 Influence of al guidelines on the external surface temperature 75
5.4.7 Guidelines on al balcony connection
77
CHAPTER 6 – CONCLUSION AND RECOMMENDATIONS
77
6.1 General conclusion 79
6.2 Recommendations
81
BIBLIOGRAPHY
85
LIST OF TABLES
86
LIST OF FIGURES

xii
Chapter 1

Introduction

1.1. Context
Today, energy efficiency in buildings is one of the prime objectives to reduce the growing energy demand
in the world and is an essential factor to reduce the energy use by 2020. Energy conservation and
renewable energy generation are two main strategies to mitigate energy demand in the world nowadays.
Although renewable energy generation has been extensively explored in past decades, the availability and
stability of renewable energy generation are still subjected to different debate in different countries. The
energy conservation unavoidably remains the most important approaches to deal with energy issues.

Building energy use including residential and commercial, contributes 20% to 40% of global energy
consumption in most developed countries [1]. The reduction in energy consumption in the building sector
can significantly reduce …. Building envelope design, including limiting the heat transfer through the
building envelope and identifying the optimal location and amount of insulation in wall section are major
strategies to improve the building energy performance.

However, even a well-designed wall that meets the thermal requirements may be subjected to
substantial energy losses as a result of the presence of thermal bridges. The presence of thermal bridges
in building envelope interfaces influences the energy consumption, durability of the building envelopes

1
Chapter 1 | Introduction

and the thermal comfort of occupants [2]. Thermal bridges are described as a part of the building
envelope where the uniform thermal resistance is significantly changed by (1) full or partial penetration
of the building envelope by materials with a higher thermal conductivity, (2) – the fabric degradation or
(3) a difference between internal and external surface area, such as occur at wall, floor and ceiling
junctions [3]. The international standard ISO 10211 establishes specific rules for the characterization of
thermal bridges in building constructions by means of numerical simulations, but these are confined to
the steady state calculations. However, to calculate the energy demand of buildings, the dynamic thermal
aspects of the building envelope and hence the thermal bridges, may be an important factor.

As the building energy demand is subjected to continuously changing outdoor and indoor conditions, the
thermal performance of the building envelope should perhaps be evaluated in terms of transient
conditions. Hence thermal bridges in building constructions should be characterized based on the
transient calculation as well.

In Northern America there is al lot of research going on about thermal bridges and building envelope
interfaces. These are, in contrast to Belgium, not standard included in the calculations for the heat losses.
Many existing and newly constructed multi-unit residential buildings (MURBs) in Canada have these
considerable thermal bridges such as balcony slabs. Hua G. investigated the influence of these thermal
bridges for a typical high-rise condominium building under various design scenarios [2]. This is one of the
subjects in the ongoing debate about thermal bridges in Canada and the implementation of them.

A valuable tool to detect thermal bridges and the related heat losses in building envelopes is the use of
infrared thermography. Recent developments in infrared detection technology have resulted in the
availability of a wide range of portable and low cost thermal cameras [4]. It is applied to provide a quick
pre-diagnostic of the entire building block and evidence of the existence of a thermal bridge in a non-
destructive way. Last couple of years, the cost of a thermal camera has been reduced significantly, and
thermal cameras become more affordable for building diagnostic purpose. A wider adoption of infrared
technology is expected in the next years. The technique visualizes the surface temperature of an object,
which is specifically of great use for building envelope inspection.

Building surface temperatures are affected by many parameters, which renders the technology primarily
fit for qualitative inspection, whereas quantitative analysis requires stringent control of boundary

2
Chapter 1 | Introduction

conditions [5]. Although infrared thermography cannot provide direct building diagnostic in current
practices, it is important to identify the optimal conditions which favor the correct application and
enables reliable interpretation of infrared thermography.

1.2. Objectives
The first objective in this study is to compare the amount of heat transfer through building envelopes
based on steady state assessment and that based on dynamic assessment. The focus is the influence of
thermal bridges and the heat losses through these building sections. The comparison between steady
state and dynamic assessment is to quantify the influence of heat losses through the building section and
hence the impact of thermal bridges can be assessed. With this research, a clear overview on the debate
in Northern America about the implementations of thermal bridges is provided. Different building
envelope interfaces are compared to each other and the application of different building materials is
analyzed.

The second objective is to investigate the impact of different boundary conditions on the results of
infrared thermography.

While in the first part the focus is to look at the total heat loss during the heating season, the aim of the
second part is to look what happens during this period. The first part emphasizes the difference in the
amount of building heat loss calculated under steady state and dynamic conditions, whereas the second
part emphasizes the heat loss process during the heating season. Different parameters affect the results
of infrared thermography, and in order to establish guidelines for the reliable application of infrared
thermography numerical simulation were executed to determine the most favorable boundary
conditions.

To achieve these goals, an overview on the heat transfer calculation and the use of different dynamic
analyses is provided. The numerical calculations of the dynamic heat losses are computed with the
thermal analysis software Voltra [6]. The second part of this work describes the use of infrared
thermography and gives an overview of the parameters that might affect the surface temperature. These
factors are investigated by means of simulations in Voltra.

3
Chapter 1 | Introduction

4
Chapter 2

Heat transfer in building components

2.1. Introduction
The energy needed to heat or cool a building can be described by the heat balance of a building. The
energy needed depends on all the heat losses and gains in the building. This heat gain or loss is at all
times equal to the sum of all the heat losses, gains and the stored heat in the building.

(2.1)

Heat gains of the sun


Internal heat gains in the building
The energy demand for heating or cooling
Transmission losses due to conduction through the building envelope
Ventilation losses as a result of convective heat transport

: Increase of the internal energy as result of heat storage in structural parts [W]

5
Chapter 2 | Heat transfer in building components

In this thesis, the focus will be on the transmission losses through the building envelope, specifically on
the impact of thermal bridges. In what follows, there is a short overview of heat transfer in buildings,
what the impact is of thermal bridges on the heat transfer through the building envelope and how to
calculate these heat losses and gains.

The first part of this chapter deals with the heat transfer of building components in steady state
conditions. This means that the heat transport through building components is calculated for constant
boundary conditions and that the properties are not changing over time. The second part of this chapter
will focus on more realistic and variable boundary conditions of the building component where the heat
flow and temperature can change over time on one or both of its boundaries. Also a short overview is
given of different methods to calculate thermal bridges in dynamic conditions. The restraints to which
building components and especially thermal bridges must comply, based on standards and codes that are
determined by governments and scientific institutions, are described in the last part of this chapter.

2.2. Thermal performance in steady state conditions

Steady state conditions means that temperatures and heat flow rates are independent of time. In this
respect, one needs constant boundary conditions, time independent material properties and constant
heat transfer. All these, but especially the constant boundary conditions, distort reality. This means that
the volumetric specific heat capacity (ρc), the thermal conductivity (λ), and heat loss are constant[7]. The
specific heat capacity will not have an impact on the steady state heat transfer.

2.2.1. HEAT TRANSFER CALCULATION

2.2.1.1 Thermal resistance of building components

A building envelope consists up of different materials, each of which can be considered as a thermally
homogeneous part or layer of the structure. Each material has its own thermal properties. How bigger
the thermal resistance value of a material layer, this is the ratio between the thickness of the layer and
the thermal conductivity, how smaller the heat flux is through this layer. The total thermal resistance of a
building component is the sum of these thermal resistance values of each material layer. The surface
resistance at the surface of the building component, is handled in part 2.2.1 of this chapter.

6
Chapter 2 | Heat transfer in building components

(2.2)
λ

R: Thermal resistance

U= : Thermal transmittance

: Heat transfer coefficient from inside and outside

: Dimension of layer I

λ: Thermal conductivity of material I

If the inside and outside temperature of the building section is known, the one dimensional heat loss
through a heterogenic construction, such as a wall, can be calculated. To implement more dimensional
heat transfer, for example in junctions between a wall and floor slab, the linear thermal transmittance
and point thermal transmittance should be considered as well. To calculate the total heat transfer
through a building section, the total thermal transmittance has to be calculated as follows:

(2.3)

: Amount of square meters of surface area loss I [m]

: Thermal transmittance of component I

: Amount of length of linear thermal transmittance j [m]

: Linear thermal transmittance of thermal bridge j

: Amount of point thermal bridges k


: Point thermal transmittance k [m]
a: Reduction factors

7
Chapter 2 | Heat transfer in building components

That formula gives an average thermal transmittance value for the related building section or building
envelope. , and are reduction factors that take into account not all parts separating the indoors
from the outside air. For those that do, the reduction factor is 1. For those separating the building from
neighboring unheated spaces, it becomes[7]:

(2.4)

: The operative temperature on the inside of the building section [K]


: The operative temperature in the neighboring unheated space [K]

In this thesis, the reduction factor is assumed 1. The total heat loss of the building section is calculated
by multiplying the total thermal transmittance with the temperature difference and the surface of the
building section.

(2.5)

2.2.1.2. Linear thermal transmittance

Thermal bridges are an important part of the building envelope that could have a big impact on the total
heat loss of buildings. They give rise to two- or three-dimensional heat flows and have a major effect on
the thermal performance of the building envelope. Compared to the two and three dimensional heat
flow, one-dimensional heat flow assumptions can cause an underestimate of the total heat loss
coefficient by about 10-40 percent for building envelopes [8]. There are two differences compared to
homogeneous building components, namely a change in heat flow rate and a change in internal surface
temperature. To consider those changes in heat flow compared to the one-dimensional heat flow, the
linear thermal transmittance is used in the two-dimensional heat flow calculations. To determine the
linear thermal transmittance ( ) for each thermal bridge there are three options [9]:

 The use of thermal bridges catalogues or handbooks where standard values are given for most
types of thermal bridges, for example in the ISO 14683.

8
Chapter 2 | Heat transfer in building components

 The use of finite element, finite difference or finite volume programs where the calculation
methodologies are more complex, but the achieved accuracy and flexibility are much higher.

The accuracy varies strongly depending on the adopted methodology. That`s why in this study, the linear
thermal transmittance is calculated from the steady-state two dimensional heat flow as a result of
simulations run by the computer software Voltra. The two-dimensional nature of the heat flow results in
heat losses that cannot be evaluated by one dimensional thermal transmittance values. The linear
thermal transmittance now defines the additional two dimensional heat loss in comparison to the one-
dimensional reference heat loss in the adjacent building components, each with their own U-value and
area [10].

(2.6)

: The linear thermal transmittance

: The global heat flow through the building section as simulated with Voltra
: The length of the junction. In this thesis, most are 1 m [ ]
: Temperature difference between the inside and the outside environments [K]

: Thermal transmittance of the adjacent building element I

: The surface within the two dimensional geometrical method for which the value
applies [ ]

The lengths that can be used to calculate the linear thermal transmittance may be measured using
internal dimensions, overall internal dimensions or external dimensions provided that the same system is
used during the calculations for all parts of a building. The system of dimensions on which the linear
thermal transmittance is based is represented by the suffix I, oi or e. These conventions are defined the
national and international standards on heat transmission. In this thesis, the one-dimensional heat flow is
determined on the basis of external dimensions, measured between the external faces of the building
components. Evidently, the convention for specifying the one-dimensional heat loss area is arbitrary,
which renders the linear thermal transmittance a non-physical value: it can be both positive as negative,
and its value is affected significantly by the geometry of the building envelope interface.

9
Chapter 2 | Heat transfer in building components

Figure 2.1 Building section with conventions for the one-dimensional


geometrical reference.

2.2.1.3. Point thermal transmittance

A point thermal bridge doesn`t have a uniform cross section along one or more of the three orthogonal
axes. It is a localized thermal bridge which influence can be represented by a point thermal transmittance
[11]. Most of the time, the influence of point thermal bridges, insofar as they result from the intersection
of linear thermal bridges, can be neglected. In this these the focus lies on the linear thermal
transmittance under transient boundary conditions, point thermal bridges will not be considered.

Figure 2.2 Building section between wall and roof with a point thermal
bridge in the corner.

10
Chapter 2 | Heat transfer in building components

2.2.2. SOLAR RADIATION


Radiation is the heat transport as a result of electromagnetical waves radiated by material surfaces. In
summer they can cause overheating of a building whereas in winter it can be an important percentage of
the heat gains. The radiation reaching the earth’s surface can be represented in a number of different
ways. The global horizontal irradiance or the total solar radiation on a horizontal surface is the total
amount of shortwave radiation received from above by a surface horizontal to the ground. The solar
radiation received from the sun that comes in a straight line from the direction of the sun at its current
position in the sky is called the direct or beam irradiance . The radiation received from the sun after its
direction has been changed is the diffuse irradiance . It does not arrive on a direct path from the sun,
but has been scattered by molecules and particles in the atmosphere and comes equally from all
directions. On a clear day, most of the solar radiation received by a horizontal surface will be direct
irradiance, whereas on a cloudy day most will be diffuse horizontal irradiance. [12]

2.2.2.1. Black body and grey body


There are two units that can be defined for the heat flow by radiation. The irradiance I is the rate power
electromagnetic radiation per unit area incident on a surface. The radiant emittance M or radiant
existence is the power per unit area radiated by a surface.

(2.7)

M: Emittance

e: Emission factor

: Black body emittance

σ= 5.67* : Stefan-Boltzmann constant

T: Absolute temperature

The total heat flow of a grey body is the sum of the heat emitted by the grey surface and the reflected
incident heat from other grey surfaces. As a result, the apparent emittance M’ of the grey area will be
larger than the emittance M:

(2.8)

11
Chapter 2 | Heat transfer in building components

: Apparent Emittance

I: Total irradiance

r= 1-e: Absorption factor

The net heat flux between the grey body and the environment is the result of the difference between
the apparent emittance and the total emittance.

(2.9)

While the radiation exchange between two grey bodies can be defined as:

(2.10)

2.2.2.2. Heat gains of the sun


To define the heat gains of the sun on a building, the irradiance of the sun is calculated. The irradiance of
a plane with slope and angle ϕ between the normal of the plane and direction of incidence of the sun
can be defined as:

(2.11)

2.2.3 SURFACE HEAT TRANSFER COEFFICIENT


The surface temperature of a structure is mainly determined by the heat transfer caused by convection
and radiation from the environment to the surface. Convection is the heat transport due to flow of fluids
and gasses. The flow of a fluid is associated with transport of sensible heat. The heat which is capacitive
stored in the fluid and causes a change in temperature can be defined as:

(2.12)

12
Chapter 2 | Heat transfer in building components

: Heat flux by convection


v: Flow rate

: Mass flow density

: Specific heat capacity

Radiation is the heat transport as a result of electromagnetically waves radiated by material surfaces. This
is discussed in the topic 2.2.2 solar radiation, of this chapter. Typically, the inside and outside
environment is defined by an air reference temperature. The heat transfer from the environment to the
surface is described by the heat transfer coefficient h. The total surface resistance depends on the
convective heat transfer coefficient and the radiative heat transfer coefficient.

(2.13)

Rs: Surface heat transfer coefficient [m2K/W]


h c: Convective coefficient
hr: Irradiative coefficient

At the external and internal surface, hc can be determined with standard values given in the 6946 ISO
standard. Although, both of them can be determined with formulas. On the external surface [13][14]in
this work the outside convective coefficient is defined as:

(2.14)

With v the wind speed [m/s]. The radiative coefficient for the internal and external surface is given by:

(2.15)

ε: The hemispherical emissivity of the surface, 0.9 is an appropriate value for internal and
external surfaces.
Σ: The Stefan-Boltzmann constant [5.67 x 10-8 W/(m2K4)]
T m: Mean thermodynamic temperature of the surface and of its surroundings [K]

13
Chapter 2 | Heat transfer in building components

In absence of specific information on the boundary conditions, design values for the surface resistance
can be used. The conventional surface resistance values Rsi and Rse for flat surfaces in contact with air are
given in the table 2.1 [15].

Surface resistance direction of the heat flow


(m²K/W) upwards horizontal downwards
0.1 0.13 0.17
0.04 0.04 0.04
Table 2.1 : Conventional surface resistance

2.3. Transient heat transfer in building components


The thermal behaviour of a building component subjected to dynamic conditions is determined by its
dynamic thermal characteristics. The component is subject to variable boundary conditions, such as
varying solar heat gains or variable air temperature on one or both side of its boundaries. Those are more
realistic conditions compared to steady state conditions, where the boundary conditions are fixed. The
properties required to compute the dynamic thermal behaviour are the dimensions of the material, the
thermal conductivity λ, the specific heat capacity c and the density ρ. The boundary conditions, surface
temperature and heat flow rate may vary periodically or non-periodically over time.

10 10
Temperature [°C]

Temperature [°C]

5 5

0 0
0 6 12 18 24 0 6 12 18 24

-5 -5

-10 -10
Time [h] Time [h]

Figure 2.3: Non-periodically boundary conditions Figure2.4: Periodically boundary conditions

14
Chapter 2 | Heat transfer in building components

2.3.1. DYNAMIC BOUNDARY CONDITIONS

2.3.1.1. Non-periodically boundary conditions


A-periodic boundary conditions occur when a building component under fixed temperatures changes to a
higher fixed temperature. These changes occur most frequently on the inside environment of a building.
The heat flux needed to raise the surface temperature of a building component is inversely proportional
to the square root of the desired warming-up time and proportional to the effusivity. How quicker a
building has to warm up, how bigger the required heat flux [16].

(2.16)

q: Heat flux needed to raise the surface temperature


: Fixed start temperature [K]
: Surface temperature [K]
t: Time needed to warm up the surface [s]

b= : The effusivity of the material

Heavy buildings that are not constantly heated, such as churches and sport halls, their warm up time and
effusivity of the inside surface do not determine the thermal transmittance factor of the building
envelope but determine the heating capacity of the building that has to be installed.

2.3.1.2. Periodically boundary conditions


During periodical boundary conditions, the parameters of the external environment are clearly marked by
a day/night fluctuation and a winter/summer fluctuation. These conditions can be simulated by using
formulas based on Fourier transformations. These describe a regular repetitive fluctuation in time and try
to approach the outside climate conditions. The use of these formulas that are explained in ISO 13786
[17] is rather complex. A more accurate method is the use of measured values of the outside
temperature. As a result of these periodic fluctuations in the external temperature, the structure will
react with a periodic response. The temperature and the heat flow density at any point in the structure is
also composed of periodic functions with the same period T as the signal of the outside temperature. The
heat flux density and the temperatures in the wall are shifted in time as a result of these varying outdoor

15
Chapter 2 | Heat transfer in building components

temperatures and specific heat capacity of the materials in the wall. The amplitude of the heat flow is
proportional to that of the outside temperature. The ratio between these two is represented by the
alternating current resistance .

(2.17)

: Alternating current resistance

: Heat flux with amplitude I

: Outside temperature with amplitude I [K]

Alternating current resistance is a building property of which the value varies according to the
characteristic period of the environmental signal. The larger the value, the better the temperature
change is muted.

(2.18)

The temperature penetration depth δ in dynamic conditions is not the same for all materials. The period
of the temperature signal and the thermal diffusivity of the material are important factors that impact the
penetration depth.

= [m] (2.19)

Figure2.5: Penetration depth

16
Chapter 2 | Heat transfer in building components

2.3.1.3. Realistic boundary conditions.


Third kind of simulation of the boundary conditions can be done by using meteorological weather data
that are measured in a weather station. Those data can be approached by formulas as described before
but the latter do not have the same accuracy.

2.3.2. THE EFFECT OF INSULATION LOCATION ON THE DYNAMIC HEAT TRANSFER


The effect of insulation location on the heat transfer characteristics of building walls and optimization of
insulation thickness is investigated in many studies [18]–[22].

The effect [23]of temperature variances on thermal inertia factors for characteristic wall configurations
have a very profound impact on the temperature fluctuations in the inner surface of building envelopes.
Al-Regib and Zubair [21] indicate that cooling loads for buildings were smaller for insulation placed on the
outdoor surface than for insulation placed on the indoor surface. Al-Sanea and Zedan [20] showed that
the insulation layer location has a significant effect on the instantaneous and daily mean loads under
transient conditions. They recommend that for spaces where the air conditioning system is switched on
and off intermittently, the insulation should be placed on the inside. The effect of wall orientation and
exterior surface solar absorptivity on time lag and decrement factor for several insulated wall
configurations was investigated by Kontoleon and Eumorfopoulou [22].It is well known that the heat
transmission load decreases without a limit with increasing insulation thickness. However, the rate of
decrease drops fast as the thickness increases. The designer should select an insulation material with the
lowest possible thermal conductivity and the highest thickness that the owner can afford. However, the
cost of insulation increases linearly with its thickness, and there is a point, for each type of insulation
material, beyond which the saving in energy use will not compensate for the extra cost of insulation
material.

Results show that yearly transmission loads are unaffected by insulation location, whereas insulation
location has a significant effect on the yearly averaged time lag and decrement factor. The maximum
temperature swings and peak load occur in the case that insulation is placed at the middle of a wall
whereas outside insulation renders the smallest fluctuation. According to Ozel [18] the best thermal
performance is obtained when insulation is placed at outside of wall [18]. This depends on the indoor and
outdoor climate.

17
Chapter 2 | Heat transfer in building components

2.3.3. DIFFERENT METHODS FOR DYNAMIC BUILDING SIMULATIONS


The dynamical behaviour of a building component can be calculated in different ways. The ISO 13786
standard [17] describes a method where the boundary conditions are considered sinusoidal. These
numerical calculations make use of matrices. The computer program Voltra calculate 3D & 2D transient
heat transfer in objects described in a rectangular grid using this energy balance technique.

Some papers suggest methodologies that try to give an analytical solution for building envelope interfaces
in transient conditions. These models aim to provide a simpler rendering than the ISO calculation, but
often lead to more complex calculations and less accurate results.

Y. Goa et. al [8] present a way to build a low-order thermal bridge additional heat loss model. It allows the
coupling with existing 1D transient thermal models, making up for the deficiency of these programs in
respect to the thermal bridge effect. Due to the equations used in this method, 1 year simulations are
impossible with complete 3D heat transfer model. The model size reduction method is used to replace
the original state model without sacrificing the main characteristics of the physical system dynamic.
Methods that are used for the model reduction are Marshall`s method, linear aggregation method,
Moore Method and thermal bridge additional heat loss reduced model which are validated by performing
an inter-model comparison. After the validation of the reduced model, it is implemented into the thermal
building simulation program, with models heat transfer trough envelope with ASHRAE 1D transfer
function approach without dynamic thermal bridge additional heat loss inclusion.

Another approach is the is the equivalent thermal wall method that has been employed for modeling the
transient response of high inertial thermal bridge [9][24]. This is a method that has been developed to
calculate an equivalent wall with the same dynamic thermal behavior as a thermal bridge. The thermal
properties of the equivalent wall are calculated using thermoelectric analogy and solving the state
equations by system identification methods.

18
Chapter 2 | Heat transfer in building components

Figure2.6: Equivalent wall method [24]

The methods described in these papers are time consuming and complicated. The models described in
these papers are not suitable for the further research in this work. The further focus is on the adequate
simulations of building envelope interfaces and the heat loss calculations during a year.

A method that is common used to calculated the heat losses on an annual basis, is the degree-days, or
degree-hour method. This method is a simple and crude method applied under static conditions. The
transient properties have been omitted in this method. Another simplification of the method is that
neither, the length nor the heating season or the temperature difference between the desired indoor
temperature and the indoor temperature without heating is dependent upon the construction design.

(2.20)

G is the number of standard degree-days or degree hours during the heating season which has typical
value for each climate. A is the indoor temperature and b is the outdoor temperature above no heating is
required anymore. The total heat loss during a heating season can then be calculated with this method by
using equation 2.21.

[J]
(2.21)

H: Heat loss coefficient

19
Chapter 2 | Heat transfer in building components

Further in this thesis, the degree hour method will be compared with numerical calculations performed
by the thermal analysis software VOLTRA[6].

2.4. Design guidelines for building envelope interfaces


There are different criteria for designing a building section. Criteria such as acoustics, air tightness, fire
resistance and durability are not discussed in this work. The focus is on the factors that can affect the
energy performance of a building. The first criterion that is discussed is the risk for mould growth and
surface condensation. Also the guidelines in respect to thermal performance of building components that
are used to design the simulated building sections are discussed in this section.

2.4.1. INFLUENCE OF MOULD AND SURFACE CONDENSATION ON BUILDING COMPONENTS

Mould growth occurs under certain conditions. It cannot develop on non-porous materials such as glass,
metal, plastic. Porous materials can retain moisture longer so they are more susceptible to mould growth.
Another parameter is the internal surface temperature that defines the risk of mould growth on the
internal surface. Last decades, a lot of research [25]–[27] was devoted to this aspect, especially mould
and thermal bridge performance. Scientific results are used in standards and technical publications with
recommendations to control and prevent the occurrence of mould growth on interior surfaces.

The risk of mould increases when the surface has a sufficiently high temperature and humidity. When
temperature decreases, the mold growth will decline and stops when it drops below 5°C. The risk for
mould growth increases if it meets these conditions over an extended period of time. For constructions, a
relative humidity of 75 percent [28] is regarded as a maximum allowable value for the structure surface.
In order to avoid mould growth on the inner surface, we can define the following boundary condition:

(2.22)

: Vapour pressure indoor air [Pa]


: Saturation value for the vapour pressure of the internal surface temperature [Pa]
: Internal surface temperature [°C]
: Relative humidity

20
Chapter 2 | Heat transfer in building components

Control of mould growth is stricter than that of surface condensation. Surface condensation occurs only
when the inner surface of a structure is colder than the dew point of the indoor air and the humidity at
the surface reaches the saturation value. To avoid surface condensation, the inside surface temperature
has to meet the requirements of equation 2.23.

(2.23)

Mould growth and surface condensation can increase when the space is inadequate ventilated or not
enough heated. These also increases when a room is not well insulated or the heat transfer from
convection and radiation is limited. The designer can only provide sufficient thermal quality in order to
avoid the risk of mold growth. The other factor are less suggestible by the designer.

The dimensionless temperature or the temperature factor f is a general factor that the internal surface
temperatures describes independent from the boundary conditions.

(2.24)

: Temperature [°C]
si, e, i: Internal surface, external, internal

To impose a requirement on the minimum temperature factor, mould growth can be prevented. If 2.22 -
2.24 are combined, a general regulation for the temperature factor can be formed as described in 2.25.

(2.25)

f: Temperature factor
: Dew point [K]

Design criteria may vary from country to country but a lower limiting value of 0,7 is accepted for the
temperature factor to reduce the risk of mould growth [16]. Sometimes, there is a suffix below the
temperature factor. This represents the surface resistance.

21
Chapter 2 | Heat transfer in building components

A well-designed detail results in a decrease of the heat loss, because of the smaller -value, and also
reduces the risk for mould resulting in a higher temperature factor f. However, a constructional detail
that meets the mould control criteria is not necessarily a detail with minimized heat flow. Thermal bridges
have often acceptable performance for mould control but with an important impact on transmission heat
loss.

2.4.2. THERMAL PERFORMANCE OF BUILDING COMPONENTS

Typically each country has its own guidelines, standards and rules in respect to the energy performance
of buildings. In this thesis, the guidelines that are determined by Flemish government in Belgium are used
to design the building sections. As a result of the Kyoto protocol signed in 1997, the European energy
performance direct for buildings was issued in 2002. This led to the EPB requirements in Flanders.

The building sections that are considered in this thesis can be classified in three categories: old building
components, building components that meet the EPB requirements and building components that meet
the passive house construction requirements. The thermal transmittance values for walls and roof that
are not in contact with the ground, can have a maximum of 0.24 W/ m2K for the normal regulations.
Passive houses have more stringent requirements: the thermal transmittance of walls and roofs should
not be higher than 0.15 W/m2K [29]. The values that are used are given in table 2.2.

22
Chapter 2 | Heat transfer in building components

Maximum allowable U-values or minimum R-values


Construction Element U max R min
[W/m²k] [m²k/W]
1 BUILDING COMPONENTS THAT SURROUNDS THE CONDITIONED SPACE
Except for the building components which form the boundary with an
adjacent conditioned space
1.1 TRANSPARENT BUILDING COMPONENTS 1.8
except for doors and gates (1.3), light facades (1.4), glass blocks (1.5) glass and
and building components which are not glass Ug,max= 1.1
1.2 OPAQUE BUILDING COMPONENTS
except for doors and gates (1.3) and light facades
1.2.1 Roofs and ceilings 0.24
1.2.2 walls not in contact with the ground, except for wall mentioned in 1.2.4 0.24
1.2.3 Wall in contact with the ground 0.40 1.5
1.2.4 Vertical and sloping building components in contact with a crawlspace
1.4
or a basement out of the conditioned space
1.2.5 Floors in contact with the outside environment 0.30
1.2.6 Other floors
0.30 1.75

1.3 DOORS AND GATES (frame included) 2.0


1.4 CURTAIN WALLS (according to prEN13947) 2.0
and
Ug,max= 1.1
1.5 GLASS BLOCKS 2.0
1.6 TRANSPARENT BUILDING COMPONENTS NOT IN GLASS 2.0
except for doors and gates (1.3) and light facades (1.4) and
Ug,max= 1.6
2 BUILDING COMPONENTS BETWEEN TWO CONDITIONED SPACES ON
1.0
ADJACENT PARCELS
3 OPAQUE BUILDING COMPONENTS WITHIN THE CONDITIONED SPACE
OR ADJACENT TO A CONDITIONED SPACE WITHIN THE SAME PROPERTY
except for doors and gates (1.3)
3.1 BETWEEN DIFFERENT UNITS
3.2 BETWEEN UNITS AND COMMON AREAS
3.3 BETWEEN UNITS AND SPACES WITH Q NON-RESIDENTIAL DESTINATION 1.0
3.4 BETWEEN SPACES WITH AN INDUSTRIAL DESTINATION AND SPACES WITH non
INDUSTRIAL DESTINATION
Table 2.2 : Maximum thermal transmittance values[29]

23
Chapter 2 | Heat transfer in building components

24
Chapter 3

Simulation model and numerical analysis

3.1. Introduction
As discussed in 2.4, there are a lot of different methods for calculating and modeling heat transfer
through thermal bridges under dynamic conditions [9][17][24]. In this work, the numerical simulations of
a thermal bridge under dynamic conditions are performed with the thermal analysis software Voltra. To
compare these results with the heat transfer in steady state calculations, the degree hour method is
used. The degree method is already discussed in 2.4. The use of the software that calculates the transient
heat transfer, according to European and international standards, is further explained in this chapter. In
the second part, the different building envelope interfaces that are used for this investigation are
presented, where after the boundary conditions that are used are discussed. In the last part, the different
results are compared and the conclusions are formulated.

3.2. Calculation of heat transfer in Voltra


VOLTRA is a thermal analysis program for transient heat transfer in three-dimensional rectangular
objects. It is a numerical method program that allows any type of thermal bridge characterization from
point of view of material properties, geometry and boundary conditions. This fact results in a wide range
of possibilities for heat transfer analysis considering steady or transient states and obtain highly accurate
values. The thermal conductivity and specific heat capacity of materials can be set as boundary conditions

25
Chapter 3 | Simulation model and numerical analysis

that may be temperature dependent. The time-dependent boundary conditions are described with
functions, either built-in functions based on parameters, or external user-defined functions based on
function values given at a fixed time interval. The thermal conductivity, specific heat capacity and
reflection factor and emissivity of a material can be described by functions that allow the change of those
properties over time. A solar processor takes into account the dynamic solar heat gains. The text output
of VOLTRA can be imported in Microsoft Excel to create graphs [6]. Further explanation of the input and
output of the program can be found in the manuals of Voltra and Trisco [30][31].

3.2.1 INPUT
The designed building envelope interfaces are discredited by means of a grid. The grid is defined in three
dimensions by gridlines. The section between the gridlines defines points that can be used to determine
the dimension of the geometry of the building component. Each block defined by these coordinates that
refer to the grid numbers of the grid can be defined by a material or surface boundary condition. The
colors window defines the colors associated with the blocks in the blocks window. Via the colors the
thermal properties related to the blocks are defined [Figure 3.1].

Figure3.1: Colours window Voltra

In this thesis, there are three different types of color defined:

MATERIAL: This defines a material with thermal conductivity, thermal density and the specific
heat capacity of this material. Also the solar reflection factor and the solar
transmission factor has to be entered. If BC_SKY defines the boundary conditions,
the emissivity of the material is also entered.

BC_SIMPLE: This defines a surface boundary condition. It is defined by a global surface heat
transfer coefficient, for combined convection and radiation

26
Chapter 3 | Simulation model and numerical analysis

BC_SKY: This also defines a surface boundary condition but with air and radiation
temperature defined separately, the radiation heat transfer at the surface is based
on view factors, and convective heat transfer is calculated separately. In both
boundary conditions, the sun can be simulated by turning the sun active status on
YES.

The properties λ, c, , h, q, a, hc, Pc, r and ρs can refer to functions that are defined in the functions
window via function reference names. When BC_SIMPL or BC_SKY is used, the parameter CEN-rule has to
be defined. It is a color subtype that defines how the thermal properties, λ or hc are calculated according
to rules approved by the Commission for European standards. In the next part model assumptions that
will be used in the further calculations are explained.

3.2.1.1. Grid
The accuracy of the numerical calculations in Voltra depends on the number of nodes that is defined by
the size of the grid. According to the standard the number of subdivisions shall be determined as follows:
the sum of the absolute values of all the heat flows entering the object is calculated twice, for n nodes
and for 2n nodes. The difference between these two results shall not exceed 1 percent. If not, further
subdivisions shall be made until this criterion is met. The use of a linear grid of 25 mm where every mesh
has the same size gives good results that satisfy the standard’s requirements [Figure 3.3]. However, Voltra
has also a special function that improves the speed of the simulations and gives more accurate results.

Figure3.2: Grid Auto split

The functions Auto split can be used to split all meshes in the specified directions into smaller parts. For
originally smaller meshes, up to a given mesh size, the mesh is split into sub-meshes having a given split
distance, and possibly a last sub mesh with width equal to the remainder after. The other larger meshes

27
Chapter 3 | Simulation model and numerical analysis

are split using a geometric subdivision using a given split ratio starting with the split distance used for
smaller meshes at both ends of the original mesh. The function “Auto Split” is intended to achieve a
reliable grid refinement using only one function call starting from the minimum grid. Further in this work,
the function auto split is used to define the number of nodes that is used during the numerical
calculations. [Figure 3.4]

Figure3.3: Linear grid of 25mm meshes Figure3.4: Auto split used a grid

3.2.1.2. Report definition


The contents of a report in the text output window, discussed in 3.2.2 output, after a calculation has been
executed, is determined by the report definition. The input of the report definition is done in the report
definition window. During the generation of a report in the Text Output window, a report line is written
per report step, which equals the calculation time step or a multiple of it, as defined by the report
frequency. The time and the properties of the material that are defined by a function, the absorbed solar
flux and surface temperature are possibilities for the output report definition. Other outputs such as the
temperature and power in one point, heat flux and temperature of a surface and the azimuth and altitude
of the sun can also be calculated.

28
Chapter 3 | Simulation model and numerical analysis

3.2.2. OUTPUT
There are two outputs in Voltra, the text output and the graphic output. A general overview is given in
the following sections for the outputs used in this work to analyse the results. More specific details of the
different outputs, especially for the graphic output, can be found in the Voltra manual.

3.2.2.1. Text output

The basic text output in Voltra reports the materials used, boundary conditions, temperature of the
material and the flow in and out of the object. More detailed results can be collected by defining the
specific results that are needed in the report definition window. After the numerical calculation, a report
is made with all the results that were asked. This report is a txt-file that can be exported to excel, where
the results are used to make graphs.

Figure3.5: Text output

3.2.2.2. Graphical output

The image given in the graphic output shows the simulated model where different properties can be
highlighted. The distribution of the surface temperature, heat flux, position of the sun and composition of
the building component are the most important ones. The graphic output is a representations of the

29
Chapter 3 | Simulation model and numerical analysis

result from the numerical calculations. Figure 3.6 shows the surface temperatures calculated by Voltra so
it look like a thermographic image, but it is only a representation of the surface temperatures.

Figure3.6: Graphical output of the surface temperature distribution

3.3. Building junctions


In what comes next, three different building envelope interfaces are discussed. Each of them is
investigated when there is no insulation present, when the building components comply with the EPB
guidelines discussed in chapter 2.4 and also when they satisfy the passive regulations. Furthermore, the
impact of the concrete slab in the roof is investigated by making a difference between a concrete slab
that doesn`t penetrate the insulation and one that does make contact with the exterior brick veneer wall.
These are schematically presented in the figures below. In order to distinguish the different building
envelope interfaces, a designation is developed.

XX#___
XX: Kind of connection: WR: Wall roof junction
B: Balcony connection
WF: Wall floor junction
# Thickness of the insulation: 14: 14 cm of insulation
___ Further specification NTB: No thermal bridge
E: Extended concrete slab
Table 3.1: Designation of the building envelope interface

30
Chapter 3 | Simulation model and numerical analysis

The cut-off planes for the three dimensional building envelope interfaces for calculating the heat flow or
surface temperatures are positioned according to ISO standard 10211 [3]. That means that the
dimensions of the cut-off planes in this study equal tare one meter. In this work, the building envelope
interfaces are presented and simulated in a simplified model. There are no mortar joints, wall ties or
other joints simulated. The rebars in de concrete slab of the balcony connections were not simulated.
Previous more detailed simulations with rebars showed that there was almost no extra heat loss
compared to the concrete slab. Figure 3.6 shows an example of a building envelope interface with the
simulated building parts.
WR24NTB

Figure3.7: example of a building envelope interface

3.3.1. WALL/ROOF JUNCTION

Flat roofs are often used in new modern buildings. A good design and understanding of the thermal
behavior is important given the increasing focus on energy efficiency. A wall/roof junction of a flat roof is
a building interface where the heat loss can be decreased by making changes in the compositions of the
sections. The impact of insulation, an extended concrete slab that makes contact with the outer leaf and
solving the thermal bridges is presented in this work.

There are a lot of compositions of wall roof junctions that exist. In this work, only the connections
between a traditional cavity brick wall and a flat roof with a concrete slab is taken into consideration to
investigate the influence of dynamic boundary conditions on the heat transfer.

31
Chapter 3 | Simulation model and numerical analysis

Wall/roof junction in buildings without insulation:


WR0 WR0E

Figure 3.8 and 3.9: Wall/roof junctions with no insulation

Wall roof junction in buildings with limited cavity wall insulation:


WR5 WR5E

Figure 3.10 and 3.11: Wall/roof junctions with 5cm retrofit cavity insulation

Wall/roof junction with insulation thickness according to EPB regulations:


WR14E WR14 WR14NTB

Figure 3.12, 3.13 and 3.14: Wall/roof junctions with 14 cm insulation

32
Chapter 3 | Simulation model and numerical analysis

Wall/roof junction with insulation according to passive house regulations


WR24 WR24NTB

Figure 3.15 and 3.16: Wall/roof junctions with 24 cm insulation

3.3.2. BALCONY CONNECTION


The presence of balconies may give rise to severe thermal bridges, which increase the heat loss and the
risk for damage due to mould. A lot of high rise residential buildings feature balconies to provide urban
dwellers with an immediate connection to the outdoors. However, the construction of these balconies
often does the same for the interior environment, providing a thermal bridge to the exterior conditions.
Thermal bridges such as these are very common in older buildings and contribute to additional heat loss
and heat gains.

Balcony connections that are common in older buildings:


B0 B5

Figure 3.17 and 3.18: Balcony connection with 0 and 5 cm insulation

33
Chapter 3 | Simulation model and numerical analysis

Balcony connections in components with insulation thickness according to the EPB regulations with and
without a thermal bridge. This according to the Umax values defined in table 2.2.
B14 B14NTB

Figure 3.19 and 3.20: Balcony connection with 14 cm insulation

Balcony connections that should be used in passive housing and buildings: (cfr supra)
B24 B24NTB

Figure 3.21 and 3.22: Balcony connection with 24 cm insulation

34
Chapter 3 | Simulation model and numerical analysis

3.3.3. WALL/FLOOR JUNCTION

A third building interface that is simulated is the connection between a wall and an intermediate floor
slab. The difference between a floor slab that is piercing the insulation or not is presented.

Wall/floor junction in older buildings


WF0 WF0E

Figure 3.23 and 3.24: Wall/floor junction with 0 cm insulation

Wall/floor junction in older buildings that have cavity wall insulation


WF5 WF5E

Figure 3.25 and 3.26: Wall/floor junction with 5 cm insulation

35
Chapter 3 | Simulation model and numerical analysis

Wall/roof junction that suffice to the EPB regulations


WF14 WF24

Figure 3.27 and 3.28: Wall/floor junction with 14 and 24 cm insulation

3.4. Model assumptions


The building envelope interfaces that are compared under steady state and dynamic conditions are
presented in previous section [3.4]. In this section, the boundary conditions and other model assumptions
that are used to simulate these building connections are discussed. These assumptions are applied to the
simulations where the heat transfer under steady state conditions is compared to that under dynamic
conditions. The assumptions for the simulations that take into account the impact of the boundary
conditions on the surface temperature, are given in chapter five.

3.4.1. BOUNDARY CONDITIONS


As input for the boundary conditions, the meteorological weather data of a reference year of Uccle,
Belgium is used. There were different time steps considered that could be used for the simulations
according to the ISO standard [17]. A time step of one day was not accurate enough, whereas the
difference in results between a time step of 10 minutes and one hour was negligible, consequently, a
time step of one hour was chosen for all the simulations.

36
Chapter 3 | Simulation model and numerical analysis

3.4.1.1. Outdoor conditions


The simulations were carried out during the period between first of October and fifteenth of April. This is
further referred to as the heating season. Between the fifteenth of April and the first of October, the
outside temperature is high enough and heating is no longer required most of the time. Also the heat
gains of the sun are higher, with the result that the heat transfer to the outside decreases. The grey area
in figure 2.9 is the period that is not taken into consideration for the calculations on the heat transfer in
this work.

30
25
20
Temperature [°C]

15
10
5
0
-5
-10
1-jan 1-feb 1-mrt 1-apr 1-mei 1-jun 1-jul 1-aug 1-sep 1-okt 1-nov 1-dec

Figure 3.29: Variable in and outdoor temperature

During the heating period, considerable variations in wind speed, sunny and overcast days, variable
outdoor air and sky temperature can be seen from figure 3.8.
30

25

20

15
Wind speed [m/s]
Temperature [°C]

10

-5

-10

-15

-20
1-okt 15-okt 29-okt 12-nov 26-nov 10-dec 24-dec 7-jan 21-jan 4-feb 18-feb 4-mrt 18-mrt 1-apr 15-apr
T outside T insdide T sky wind speed
Figure 3.30: Variable outdoor climate date for the heating season, Uccle, Belgium and the variable indoor temperate regimes
measured in a dwelling of a Belgian family

37
Chapter 3 | Simulation model and numerical analysis

3.4.1.2. Indoor conditions


In most of the simulations, the indoor temperature is described by data measured in a Belgian dwelling
[Figure 3.7] that is representative of typical interior conditions. This should give a good approximation of
the fluctuation temperatures in reality. To assess the effect of different materials on the heat storage in
materials, simulations are also done with a constant temperature of 20 degrees Celsius.

3.4.2. Calculation method


To show the calculation method that is used, the building envelope interface WR14 is given as an
example. The same method is used for the rest of the building junctions.

3.4.2.1. Steady state


First the thermal transmittance of the wall and roof is calculated. Thereafter, the two dimensional heat
loss through the building envelope is calculated with the software Voltra to determine the linear thermal
transmittance factor. Not all the sun gains are useful, and a detailed analysis of these gains should be
investigated on basis of an entire building. Because only building envelope interfaces are simulated, the
influence of the sun on the heat loss between steady state and dynamic is out the scope of this thesis.
During this simulation, the boundary conditions are simulated as BC_SIMPLE with outdoor temperature of
0 degrees, indoor temperature of 20 degrees, surface resistance factor according to table 2.1 and no
influence of the sun. This approach complies with the calculation guidelines for the Flemish Energy
Performance Directive.

The second step is to calculate the total heat loss during the heating season. This is based on the degree
hour method. The amount of degree hours is the sum of hourly temperature differences between indoor
and outdoor temperature. During the heating season and according to the used weather data, there are
70313 degree hours. The total heat loss through the building envelope interface is calculated with eq. 3.1.

(3.1)

In this work, degree hours are used instead of degree days and these represent the temperature
difference between indoor and outdoor while the degree days explained in chapter two are based on the
assumption that there is no need of heating anymore with a certain external temperature. The degree

38
Chapter 3 | Simulation model and numerical analysis

day method is used to calculate the energy loss in buildings while the difference between the internal
and external temperature each hour makes a simplification possible between steady state and transient
conditions.

Another consideration that is taken into account is the percentage of the total energy loss that is lost
trough the junction. This heat loss can be compared between the steady state calculations and the
dynamic calculations. Figure 3.31 represents an example of the calculations to determine the total heat
loss in steady state conditions and to observe the heat loss trough the junctions.

Figure3.31: WR14 building envelope interface

U value wall= = 0.217 W/m2K

U value roof= = 0.237 W/m2K

- Value= = 0.152 W/mK

Qwall= = 20.006 kWh

Qroof= = 23.484 kWh

Qjunction= = 10.685 kWh

Qtotal= = 54.176 kWh

39
Chapter 3 | Simulation model and numerical analysis

3.4.2.2. Dynamic
The total heat loss in dynamic situation is calculated by Voltra. There is a differentiation between the heat
loss from the inside to the object and from the object to the outside. Only the heat losses from the inside
to the building component are taken into consideration. This is because the energy that is needed in the
building is the energy that is lost into the construction. The heat exchange between the outdoor
environment and the building component is not useful for these calculations. The heat transfer between
indoor and the construction can be divided into heat losses, the heat that is lost into the construction,
and the heat gains. This is the heat that the construction gives back to the indoor environment owing to
the fluctuating temperature and the heat capacity of the construction.

Figure 3.32: Heat flow in and out of the object

When the indoor temperature is constant, there is no heat transfer back to the indoor environment.
When the simulations are being done with fluctuating indoor temperature, the difference between the
flow in and flow out is used as the total heat transfer. As shown in figure 3.11, there is no heat flow out of
the object into the indoor environment when the indoor temperature is constant. When the indoor
temperature is fluctuating, the heat loss into the object is also fluctuating. The flow out of the object
happens at night.

40
Chapter 3 | Simulation model and numerical analysis

40

30
Temperature [C]
Heat flow [W]

20

10

-10
1/5/14 0:00 1/6/14 0:00 1/7/14 0:00 1/8/14 0:00 1/9/14 0:00 1/10/14 0:001/11/14 0:001/12/14 0:00
T outside T inside Flow out Flow in Flow in; T cte

Figure 3.33: Heat flow in and out of the building envelope interface

3.4.2.3. Comparison
The total heat loss that can be considered during the heating season under dynamic conditions for the
building envelope interface WR14 is equal to 53,751 kWh. To compare how much energy is lost through
the different building components, each building component is separately simulated in Voltra, where
after the total heat loss of each building component is subtracted from the total heat loss of building
envelope interface. This results in the heat loss of the junction in dynamic conditions. To compare the
different results, heat loss in steady state and dynamic conditions are presented in percentages.

Steady state conditions Dynamic conditions Ratio: dynamic/steady


Total 54.176 kWh 53.751 kWh 99%
Wall 20.006 kWh 19.814 kWh 99%
Roof 23.484 kWh 23.192 kWh 98%
Junction 10.685 kWh 10.746 kWh 101%
Table 3.2: Comparison of the results of the building envelope interface WR14

If the dynamic results are compared to the steady state, there can be determined that there is more heat
loss in transient conditions over the whole building envelope interface than calculated in steady state
conditions. Although in the junctions occurs more heat loss in transient than in steady state conditions.
This is because by the thermal inertia that is a part of the dynamic calculations and caused by the lower
thermal inertia of the building junction compared to the wall and roof.

41
Chapter 3 | Simulation model and numerical analysis

3.5. Comparison of the steady state and the dynamic heat loss calculations
In the next paragraph, the different results of al the building envelope interfaces are presented. First, the
total heat loss in steady state and dynamic is shown were after the percentage between both are given.
Thereafter, the percentage through the junctions and connections in steady state and transient
conditions are compared by expressing the ratio between the total heat losses.

A) WALL/ROOF JUNCTIONS
Total heat loss Junction
Steady state Dynamic Ratio Steady state Dynamic
WR0E 315,3182 kWh 305,451 kWh 97% -14,69 % -17,59 %
WR0 310,457 kWh 300,731 kWh 97% -16,82 % -19,78 %
WR5E 252,951 kWh 252,446 kWh 99% -9,69 % -9,61 %
WR5 237,798 kWh 237,305 kWh 99% -16,68 % -16,6 %
WR14E 83,004 kWh 82,55 kWh 99% 47,6 % 47,9 %
WR14 54,176 kWh 53,751 kWh 99% 19,72 % 19,99 %
WR14NTB 42,504 kWh 41,735 kWh 98% -2,32 % -5,57 %
WR24 36,492 kWh 36,073 kWh 98% 22,49 % 22,9 %
WR24NTB 27,106 kWh 26,7254 kWh 98% -4,36 % -4,06 %
Table 3.3: Total heat loss wall/roof junction and the percentage through the junction in steady state and dynamic

B) BALCONY CONNECTIONS
Total heat loss Junction
Steady state Dynamic Ratio Steady state Dynamic
B0 264,974 kWh 264,395 kWh 99% 11,78 % 13,04 %
B5 134,684 kWh 134,083 kWh 99% 40,58 % 40,84 %
B14 82,231 kWh 81,688 kWh 99% 58,96 % 59,08 %
B14NTB 41,731 kWh 41,253 kWh 98% 19,12 % 18,97 %
B24 62,93 kWh 62,378 kWh 99% 66,91 % 67,04 %
B24NTB 28,828 kWh 28,369 kWh 98% 27,76 % 27,52 %
Table 3.4: Total heat loss balcony connection and the percentage through the connection in steady state and dynamic

42
Chapter 3 | Simulation model and numerical analysis

C) WALL/FLOOR JUNCTIONS
Total heat loss Junction
Steady state Dynamic Ratio Steady state Dynamic
F0E 257,029 kWh 256,457 kWh 99% 9,05 % 10,34 %
F0 244,408 kWh 243,835 kWh 99% 4,35 % 5,7 %
F5E 124,489 kWh 123,943 kWh 99% 35,72 % 36,01 %
F5 81,07 kWh 81,08 kWh 100% 1,29 % 2,17 %
F14 33,926 kWh 33,474 kWh 98% 0,52 % 0,14 %
Table 3.5: Total heat loss wall/floor junction and the percentage through the junction in steady state and dynamic

If the total heat losses are compared between the steady state and the dynamic calculations, there is
almost no difference. The difference in total heat loss is devoted to the fact of thermal inertia of the
construction.
The percentages of heat loss in the junctions and connections represent actually the linear thermal
transmittance. As shown in table 3.3, some of these values are negative, other are positive. How more
negative the percentage of heat flow the better the thermal performance of the junctions is compared to
the thermal performance of the wall and roof. Table 3.4 represents the result of heat losses through
balconies, these percentages are a lot higher than the other examples because the concrete slab has an
bigger impact on the total performance of the junctions.

3.6. Conclusions
The main findings after comparing these results is that there are almost no differences in total heat loss
between dynamic and steady state calculations as shown in paragraph 3.5. But the heat loss through the
junctions or connections, these are the places that are the most sensitive to heat losses compared to the
rest of the building components, is greater under dynamic situations than calculate in steady state
situations. More attention should be paid during the heat transfer and the effect of faze transition and
not to the total heat loss during one year.

Although there is a little difference between the results in steady state and transient conditions, this is
neglectable. Out of these result can be concluded that the linear thermal transmittance calculated under
steady state conditions is a very good approximation of the two dimensional heat losses through a

43
Chapter 3 | Simulation model and numerical analysis

building envelope in transient conditions. This method can be used to calculated the heat losses through
the building junctions that are presented and simulated in the paper of Dr. H. Ge.[2]. In general, the linear
thermal transmittance can be the answer on the ongoing debate in Northern America, where heat losses
caused by junctions, connections and thermal bridges are not standard concluded in the calculations of
heat loss in a building.

44
Chapter 4

Infrared thermography
4.1. Introduction
The total heat loss of dynamic calculations and the influence of thermal bridges on these heat losses are
presented in the previous chapter. In this chapter, the methodology to detect thermal bridges by means
of infrared thermography is presented.

Infrared thermography has been used in many disciplines such as medical care, electronics and food
industries. Recent developments in infrared detection technology have resulted in the availability of a
wide range of portable and low cost thermal cameras for the building industry [4]. In current practices,
thermal infrared imaging is commonly used for diagnosing building envelope problems. It provides a way
to detect the location and the process of heat transfer in a building envelope. It is a non-destructive
testing method to collect data that is used for investigating the air-tightness of buildings or the detection
of thermal bridges in the building envelope. The technique instantly visualizes the surface temperature of
a whole building part. Diagnosing building problems by using infrared thermography requires knowledge
of the building structure and the mechanism of heat, air and moisture (HAM) transport through the
building envelope. However, many factors that affect thermography imaging need to be taken into
consideration to ensure a reliable analysis and conclusions. Some parameters can have a major influence
on the surface temperature but are often disregarded and may lead to false conclusions. Many factors
influence the mechanism of HAM transport, such as thermal conductivity, effusivity, emissivity,

45
Chapter 4 | Infrared thermography

absorption, reflection, vented cavities, failures of building systems and heating patterns. These factors
have an impact on the final judgment of thermography imaging results and lead to distorted conclusions.
Infrared thermography can be a powerful tool for fast and accurate building diagnostics. It reveals a
different dimension and enables an auditor to substantiate observations when there are no obvious
problems based on visual inspection. This can reduce guesswork and identify abnormalities that could
otherwise go undetected before they evolve to more serious problems or even costly failures and
widespread damages [5]. It facilitates more solid hypotheses that can be deduced from thermographic
images. Potential future failures and serious damages can be minimized. Moreover, thermography is
simply the only technique that allows a rapid inspection of surface temperatures. The proper use of
infrared imaging and the influencing factors are discussed further in this chapter.

4.2. Use of infrared thermography

An infrared camera is used to measure the surface temperature of an object, based on the emitted
radiant thermal energy from its surface. The energy received by an infrared sensor of the thermal camera
[figure 4.1] consists of the energy emitted from the target and its surroundings. Furthermore, infrared
emission of the atmosphere between the camera and the object contributes to the long wave radiation
from behind the element and is partially transmitted. The air/atmosphere (number 3 on the picture)
between the camera and the object, and in some cases irradiation, ca have an impact, but this is less
important. Number one on the picture represents the reflection of the surrounding irradiation on the
object. These factors should be taken into consideration for evaluation of thermal bridges.

1:Surroundings 2:Object 3:atmosphere 4:camera sensor


Figure 4.1: Schematization of the heat transfer process between infrared camera and the object.

46
Chapter 4 | Infrared thermography

The standard equation that represents the radiation that is received at the sensor of a camera can be
described as:

W: Irradiance
ε: Emissivity of the surface
: Transmittance
0
: Black body

Infrared thermography imaging on building envelopes is mainly used for detection of thermal bridges.
Besides the heat losses through thermal bridges, the risk on mold is another negative aspect. The risk on
mold can be evaluated by determined the temperature value. The incoming energy to the infrared
camera is converted to temperature values according to the Stefan-Boltzmann law. When the parameters
for atmosphere and surrounding temperature and emissivity are correctly set in the camera software,
thermography allows the user to determine the instantaneous surface temperature of the building
envelope. Internal and external air temperature can easily be measured on site so that the temperature
factor can be calculated. The temperature factors of the building envelope interfaces are used to classify
the type of thermal bridges. However, the infrared image is only an instantaneous photograph during
transient conditions. A general overview of existing guidelines to minimize the error on the dynamic
temperature factors, with special attention for building envelope interfaces are explained in the next
section.

4.2.1. ACCURACY OF INFRARED PICTURES


The infrared camera performs better as a relative temperature measurement device than as an absolute
temperature measurement device [32].The accuracy of an infrared camera is ± 2.0 °C. However, in
controlled indoor conditions with excellent knowledge of the boundary conditions, temperatures can be
measured with an accuracy of ± 0.5°C [33]. Therefore, the accuracy for outdoor conditions is between 1
and 2 °C, whereas indoor under controlled conditions, the error can be reduced to ± 0.5°C. As shown in
picture 4.2, a temperature difference is already clearly evident if there is a difference of 0.1°C. [figure 4.2
and 4.3] As a result of this, the detection of thermal bridges is not complicated. The estimation of the
heat loss through the thermal bridges is more difficult to determine.

47
Chapter 4 | Infrared thermography

Figure 4.2: IR picture with scale between 13.8°C and 17.5°C

Figure 4.3: IR picture with more detailed scale between 17°C and 17.7°C

4.3. Influence factors on infrared thermography

4.3.1. GENERAL
During a thermographic survey, there are a lot of parameters affecting the results. Not only the
environmental factors, but also material properties, specifications of the thermographic camera and
location-related parameters are important. Although a lot of papers and standards provide guidelines for
the boundary conditions, there are a lot of discrepancies between these guidelines. Most of these papers
do not distinguish between heavy walls and light walls or indoor and outdoor measurements while they

48
Chapter 4 | Infrared thermography

should be a difference because thermal inertia has an influence on the accuracy of the results in dynamic
conditions.

In the next section, the focus is on the environmental parameters because the effect of these factors can
be simulated using the computer software Voltra. Material properties such as emissivity, reflectance,
location-related parameters and the specifications of the thermo graphic camera are not within the scope
of this study.

4.3.2. ENVIRONMENTAL FACTORS


There are a lot of environmental factors posing an impact on infrared imaging such as the reflecting
temperature [34], wind speed [4][5] [35], solar radiation [4], temperature gradient [4], temperature
difference between indoor and outdoor [35], intensity and direction of rainfall [5], irradiation to the sky
dome [4][5] and the transmissivity of the atmosphere [36]. In this study, the focus is on the wind speed,
solar radiation and the temperature difference. These are taken into consideration during the numerical
simulations in Voltra.

4.3.2.1. Wind speed


The wind speed influences the result of outdoor measurements more significantly than that of indoor
measurements. When the influence of the wind is investigated for a measurement of an exterior surface
it is quite obvious that it will have a significant impact.

The convective heat transfer is related to the wind speed. A high wind speed on the outside surface can
result in more cooling or heating, depending on the temperatures on both sides of the constructions.
Accuracy of the internal measurements are less depended on the wind speed, although heat losses can
occur due to air leakage in the thermal insulation and this can be observed by a thermographic camera.

The cooling or heating of external surface temperatures are increased with the presence of wind, the
temperature differences between the thermal bridges and the adjoining components based on
measurement from the interior are not the same as those based on measurements from the exterior.
In papers, the guidelines that are provided in respect to the wind speed are typically derived for
measurements on the outside. During the simulations in this work, a difference is made between indoor
and outdoor applications of the infrared camera. General limitations of the wind speed that can be

49
Chapter 4 | Infrared thermography

allowed during an infrared survey differ significantly from author to author. The table below shows the
principal values of the maximum wind speed for each paper. These are the values that they propose as a
maximum wind speed where infrared imaging can be applied. Higher then these values, according to
those papers, it is not reliable anymore.

5 m/s [5][35]
2 m/s [4]
Table 4.1: limitations for the wind speed

4.3.2.2. Temperature difference

The temperature difference between the indoor and outdoor environment also plays an important role in
the performance of thermal imaging. The greater the temperature difference between the indoor and
outdoor environment, the larger the heat flux passing through the building envelope. This results in the
more obvious images of the thermal bridges that can be noticed by a thermographic camera, as there is a
larger temperature gradient over a surface because differences in thermal properties become more
pronounced. Again, a lot of different values were found in the literature. Some papers differentiated
between measuring insulation errors and air leakage measurements. In this study, the focus is on the
insulation defects and thermal bridges. Most common recommendations are shown in table 4.2.

ΔT > 10°C [37][35]


ΔT > 10°C – 15°C [38]
Table 4.2: limitations for the temperature difference

4.3.2.3. Solar radiation


The thermal inertia of an object is very important when the influence of solar radiation is considered.
Most papers recommend that direct sunlight should be avoided during the thermographic measurement.
Direct solar radiation increases the surface temperature of the façade, and therefore renders a distorted
picture of the surface temperature of the building component. Most of the authors do not distinguish in
guidelines for the different types of thermal mass in building components. Lehmann [4] distinguishes in
his paper three kinds of constructions: those with small thermal inertia, those with normal thermal
inertia and the heavy constructions with large thermal inertia. The guidelines are shown in table 4.3.

50
Chapter 4 | Infrared thermography

Small thermal inertia Normal thermal inertia Large thermal inertia


Thermographic investigation Thermographic investigation Thermographic investigation
possible a couple of hours after possible 24 hours after the possible 48 hours after the
the direct sun light [4]. direct sun light [4]. direct sun light [4].
Table 4.3: limitations for the solar radiation

4.4. Conclusions
It can be concluded that for measurements from the interior side the temperature difference and solar
radiation are the most significant factors, whereas wind effects only become important for
measurements from the exterior side. Some of the recommended guidelines are tested in chapter five
even as development of new ones.

51
Chapter 4 | Infrared thermography

52
Chapter 5

Development of guidelines to detect different


building envelope interfaces
5.1. Introduction
The influence of several parameters on the result of an infrared thermography survey [4][5][35][37][38]is
already discussed in chapter four. In this chapter the influence of three factors, wind speed, solar
radiation and temperature difference, are analyzed by using numerical calculations. Eight building
junctions and six balcony connections are modelled with the thermal analysis software Voltra to
investigate and analyze the influence of the three boundary conditions.

Two sequences are involved in the simulation process. The aim of the first sequence is to analyze the
impact of wind speed, solar radiation and temperature difference on the inside surface temperature
whereas the second sequence focuses on the influence of the outside surface temperature. The first part
involves the simulation of eight wall/roof connections. The compositions of these connections are
explained in chapter 3. Thereafter, the junction WR3 is selected as a reference case to generalize the
guidelines for applying indoor infrared thermography. Then, general guidelines are developed to detect
different junctions by infrared imaging and these are applied on all wall/roof junctions to verify them.

53
Chapter 5| Development of guidelines to detect different building envelope interfaces

The second set of simulations focus on six types of balcony connections. First, the guidelines for internal
infrared thermography are applied on these balcony junctions. Thereafter, the connection B5 is selected
as a reference case which is adopted to develop the guidelines for applying outdoor infrared
thermography. The guidelines are further verified to the rest of the five balcony connections. Criteria are
developed to classify the appropriate conditions for accurate infrared camera survey to analyze the
difference between the balcony connections.

5.2. Model assumptions

5.2.1. GENERAL
The numerical simulations of the different building connections mentioned in section 3.1 are performed
in Voltra. The aim of these simulations is to assess the external surface temperatures se and the internal
surface temperatures si of building connections from which the temperature factor can be calculated.
The temperatures are measured at different points as shown in figure 5.1 and 5.2. The points one to five
are specified as internal surface temperatures whereas points six to eleven are designated as external
surface temperatures. The temperature factor is dimensionless in order to compare the values of
temperature factors obtained in steady state and to those obtained in dynamic conditions. The influence
of the boundary conditions on the error of the dynamic versus the steady state temperature factors is
further investigated.

Figure 5.1 Balcony connection with surface temperatures Figure 5.2 Wall/roof connection with surface temperatures

54
Chapter 5| Development of guidelines to detect different building envelope interfaces

5.2.2. BOUNDARY CONDITIONS


The use of the computer software is explained in chapter three. During these simulations, the boundary
conditions are simulated in a more realistic way than in chapter three. The weather data of a reference
year in Uccle, Belgium is used as an input parameter. The difference between these simulations in this
chapter and the ones in chapter three is the boundary condition setting. BC_SIMPL is used in chapter
three whereas BC_SKY is used in this chapter. BC_SKY is used because it gives a more detailed
representation of the boundary conditions and takes into consideration solar gains. [chapter 3.2]

The convective heat transfer coefficient is determined by the wind speed in transient state and also the
solar radiation is taken into consideration in this chapter. These boundary conditions, the outdoor air
temperature and radiation temperate are determined by the weather data from Uccle. The indoor
temperatures are obtained from measurements in a Belgian dwelling.

The temperature factor under steady state conditions is calculated in Voltra with a constant external
temperature of 0 degrees Celsius and a constant internal temperature of 20 degrees Celsius. The surface
heat transfer coefficient is determined according to table 2.2.

5.2.3. CALCULATION METHOD


The aim of this section is to prioritize the significance of the outdoor parameters: wind speed,
temperature difference and solar radiation on the surface temperature. The influence of these
parameters is investigated in Voltra, and the results are used to assess to what extent the severity of
specific thermal bridges would be evaluated correctly if a thermographic camera were to be used. The
surface temperatures are then converted to temperature factors which are dimensionless.

For each time step of the transient simulations, the surface temperature is reported and the temperature
factor is calculated. These dynamic temperature factors are compared to those calculated in steady state
conditions for a specific location. These results are presented as box plot distribution shown in figure 5.3.
It shows the minimum and maximum temperature factors throughout the heating season at a particular
point. The box indicates three fractals: 95%, median and 5% represent upper, middle and lower lines
respectively. The black point represents the temperature factor measured in steady state conditions. The
aim of this analysis is to evaluate whether or not thermography can be used to detect thermal bridges
(qualitative assessment), or even to get a rough estimate of the temperature factor (quantitative

55
Chapter 5| Development of guidelines to detect different building envelope interfaces

assessment). Since the temperature factor calculated under static boundary conditions is often used to
assess the risk for mould growth, a reliable estimate using thermography would be a great asset for
planning building retrofits. Consequently, each time step of the dynamic simulations is used to calculate a
temperature factor, as if a separate thermographic inspection is done each time step. The temperature
factor is calculated from the instantaneous surface temperatures. By analyzing the impact of the
boundary conditions and defining restrictions with respect to parameters such as temperatures
differential, it might be possible to get a more reliable estimate. Furthermore, in performing
thermographic inspections, the building auditor can only rely on one measurement, and does not possess
the overall spread of the results over time. The box plot can thus be understood as an uncertainty interval
that can be used to determine the overall spread of the results over time. For hygrothermal risk
assessment a higher temperature factor is preferred, therefore a conservative approach would be to
subtract a confidence interval from the temperature factor derived from one single thermographic
measurement. The difference between the steady state temperature factor and the 5% fractal can thus
be read as a 95%-reliability interval that can be used by building auditors.

Figure 5.3: Box plot distribution

The first step is to present the temperature factor for certain points within the building components.
These points are situated in the corners of the junctions, where the lowest temperature factor can be
expected. Next to that, the difference in the temperature factor between the point in the junction and

56
Chapter 5| Development of guidelines to detect different building envelope interfaces

one measured on the surface 1m away from the junction can be used to evaluate to what extent a
thermal bridge is evident under different boundary conditions. After that, one of the building envelope
interfaces is chosen as an example to further explain the influence of the boundary conditions. This is
done by the infrared imaging on wall/roof junctions and for the balcony connections.

5.3. Wall/roof junction

5.3.1. GENERAL
Before the influence of these environmental parameters is presented, the difference between the
temperature factor in steady state and the average one calculated in dynamic are compared as shown in
table 5.1. It is known that the better the thermal performance of the wall/roof junction is, the smaller the
difference is between the two factors. The difference between those two factors is caused by the impact
of the sun and thermal inertia of the materials in transient conditions. These have not been included in
the steady state calculations. The results are compared to the steady temperature factor, but the aim is
to decrease the range between the 5% fractal and the 95% fractal so the accuracy of the rough
estimation of the temperature factor is as high as possible. An overview of the f-factors is given is table
5.1. First the temperature values measured in point three are given were after the difference between
the f-factor in point one and point three is given.

F3 F1-F3
Steady state Dynamic Difference Steady state Dynamic Difference
WR0E 0.570 0.536 0.034 0.237 0.255 0.018
WR0 0.596 0.557 0.039 0.210 0.233 0.023
WR5E 0.602 0.560 0.042 0.331 0.348 0.017
WR5 0.680 0.623 0.057 0.253 0.284 0.031
WR14E 0.740 0.713 0.027 0.234 0.250 0.016
WR14 0.860 0.842 0.018 0.106 0.123 0.017
WR14NTB 0.914 0.900 0.014 0.054 0.065 0.011
WR24 0.899 0.884 0.015 0.081 0.095 0.014
WR24NTB 0.941 0.933 0.008 0.040 0.046 0.006
Table 5.1: f-Factor of wall/roof junction in steady state and dynamic conditions

The same conclusion as in chapter three can be concluded. The higher the thermal performance of the
building envelope interface, the smaller the difference between the temperature factor in steady state

57
Chapter 5| Development of guidelines to detect different building envelope interfaces

and transient conditions. This is mainly caused by the sun. The sun has a bigger impact on the internal
surface temperature of building envelopes that are not insulated than on well insulated building
envelopes. In figure 5.4 and 5.5 the distribution is provided of the f factor calculated respectively in point
three and the difference between point one and point three of all the wall/roof junctions discussed in
chapter three. On the x-axis, the name of each building envelope interface is represented according to
table 3.1.

1,50 1,35
1,28 1,25
1,30 1,16
1,05 1,03 1,02
1,10 0,95 0,95

0,90 0,91 0,90 0,94


0,86
0,70 0,74
0,68
0,57 0,60 0,60
0,50 0,61
Temperature factor

0,53
0,30 0,41
0,10
0,18
-0,10

-0,30

-0,50

-0,70
-0,61
-0,90
-0,83
-1,10 -1,00 -0,96
-1,03
-1,30
WR0E WR0 WR5E WR5 WR14E WR14 WR14NTB WR24 WR24NTB

Figure 5.4: f-Factor distribution of the different wall/roof junctions in point 3

The temperature factors are widely spread throughout the heating season. If you compare the
distributions between the different wall/roof junctions, it is not possible to separate the different building
junctions. There can be made a distinction between the first four junctions with low thermal
performances and the last four with higher thermal performances but in order to make further distinction
between the junctions, general guidelines should be applied. The result that can be measured for a
WR14E junction could easily be a result of the WR0E junction. A range of restrictions for boundary
conditions during an infrared thermography survey may provide more reliable estimates.

58
Chapter 5| Development of guidelines to detect different building envelope interfaces

1,60
1,43
1,39
1,40 1,30

1,20
1,05
1,01
1,00
Temperature factor

0,80
0,60
0,60
0,45

0,40
0,33 0,24
0,24 0,25 0,17
0,20 0,21 0,23
0,11 0,08
0,05 0,04
0,00
0,05 0,06 0,06 0,06 0,04
-0,01 0,02
-0,02 -0,04
-0,20
WR0E WR0 WR5E WR5 WR14E WR14 WR14NTB WR24 WR24NTB

Figure 5.5: F1 minus F3 distribution of the different wall/roof junctions

Even if the distribution of the difference between the two temperature factors is considered,
straightforward conclusions are not self-evident when an infrared image is taken when no restrictions are
applied. The difference between point 1 and point 3 can be used to determine the difference between a
wall/roof junction where the concrete roof slab is in contact with the exterior brick veneer wall, and a
situation where cavity or insulation in the wall is continuous.

The building envelope interfaces with a lower thermal performance, or in case of WR5E a big thermal
bridge, have a wider spread of values. Whereas the dynamic temperature factors of building junctions
that suffice to the regulations and have a better thermal performance, have less distributed values. To
investigate the guidelines given in chapter four, further tests are performed on junction WR5E. Because
during the infrared survey of a building envelope, the composition is not always known, the worst case
scenario is taken as an example.

5.3.2. INFLUENCE OF THE WIND SPEED ON THE INTERNAL TEMPERATURE FACTOR.


The first simulations on the WR5E junctions are described in figure 5.6. In each simulation, one of the
boundary conditions is set as a constant to investigate which boundary conditions have an influence on
the internal surface. The temperature factors are calculated for point three of the internal building
envelope interface. The temperature factor in steady state for this junction is equal to 0.602.

59
Chapter 5| Development of guidelines to detect different building envelope interfaces

1,40
1,20 1,09 1,07
1,02
0,93
1,00
0,74
0,80
0,60
0,40
0,43
0,20
Temperature factor

0,00
-0,20
-0,40
-0,60
-0,80
-1,00
-0,96 -0,92 -0,94
-1,20
-1,40
-1,60
-1,80
-1,69
-2,00
A B C D E

Figure 5.6: WR5E f-factor distribution different parameters

A: The indoor surface temperatures in point three during the simulations with the realistic weather data.
B: The indoor surface temperatures in point three during the simulations with a constant indoor
temperature of 20 degrees.
C: The indoor surface temperatures in point three during the simulations with a constant indoor
temperature of 20 degrees and a constant wind velocity of 2m/s.
D: The indoor surface temperatures in point three during the simulations with a constant indoor
temperature of 20 degrees but simulated without the solar irradiation.
E: The indoor surface temperatures in point three during the simulations with a constant temperature
difference of 15 degrees.

There is almost no difference between the results with a constant temperature factor of 20°C or a
fluctuating internal temperature. The results of box plot B compared to the results with a constant wind
speed [box plot C], show that the wind has little influence on the temperature factors on the internal
surface. The 5% and 95% fractal increase only 0.02 compared to the original situation. Box plot D and E
show that the main influence factors on the internal temperature factor are the sun and the temperature
difference between inside and outside.

60
Chapter 5| Development of guidelines to detect different building envelope interfaces

5.3.3. INFLUENCE OF THE TEMPERATURE DIFFERENCE BETWEEN INDOOR AND OUTDOOR ON THE INTERNAL
TEMPERATURE FACTOR

To determine the influence of the temperature difference on the accuracy of thermal imaging, a
simulation on the WR5E junction is performed where no sun is taken into account and the wind is
constant. The temperature factors measured each hour during the simulation period are shown in figure
5.7 against the temperature difference.

0,8
0,7
0,6
0,5
0,4
0,3
0,2
Temperature factor

0,1
0
-0,1
-0,2
-0,3 Steady state f factor

-0,4
-0,5
Dynamic f factor
-0,6
-0,7
-0,8
-0,9
0 5 10 15 20 25 30
Temperature difference [°C]
Figure 5.7: All dynamic temperature factors against the temperature difference

The results show that infrared imaging with a temperature difference between internal and external
environment less than 10 °C is not reliable. This is because of the large dispersion of the values that are
measured for a temperature difference lower than 10 degrees, as shown on figure 5.7. There are
temperature factors that are lower than zero. In a second graph [figure 5.8] the temperature factors are
presented in a box plot with a distribution according to figure 5.3 for each degree of temperature
difference, starting from 9 °C. It is clear that the dispersion of the f-factors lower than 10 degrees is
exponential increasing.

61
Chapter 5| Development of guidelines to detect different building envelope interfaces

0,75 0,74

0,70 0,66 0,65 0,66 0,66


0,64 0,64 0,63 0,63 0,64 0,64 0,63
0,65
0,60
0,55
Temperature difference

0,50
0,45
0,47 0,48
0,40 0,44
0,43
0,35
0,36 0,37
0,30
0,25 0,30
0,20 0,23
0,20 0,21
0,15
0,16
0,10
0,05 0,10
9-10 10-11 11-12 12-13 13-14 14-15 15-16 16-17 17-18 18-19 19-20 20-30
Temperature difference within 1°C [°C]
0
Figure 5.8: Temperature factor distribution within 1 C temperature difference bin

In this case, the focus is on the range between the 5% fractal and the 95% fractal. The smaller the spread
of the temperature factors this are, the smaller the deviation of the temperature factor is compared to
the f- factor in steady state. The temperature factor in steady state represents the risk of mold. Also how
smaller to the spread of the temperature factors are and the more accurately the infrared image
represents the approach of the temperature factor. A minimum temperature difference of 10°C degrees
is still very optimistic. The distribution of the values becomes more acceptable once there is a
temperature difference of 15 degrees as show on figure 5.7 and 5.8. For the example of the WR5E
junction, 90% of the f-factors that are calculated for a temperature difference higher than 15 °C are in a
range of 0.16. While for temperature factors measured with a temperature difference higher than 10 °C,
the range is 0.27. A more stringent guideline is that an infrared survey should not be performed when the
temperature difference is lower than 18 degrees. The range of the f-factors between the 5% and the 95%
fractal is only 0.1. It is not useful to define even more stringent guidelines for the temperature difference.
The range will be almost the same and the periods during the year where you can do an infrared
thermography survey will decrease exponentially. The guidelines for the temperature difference are
presented in table 5.2. These are provisional guidelines that will be combined with the solar guidelines
later in this chapter.

62
Chapter 5| Development of guidelines to detect different building envelope interfaces

Minimum requirements Recommended requirements Stringent requirements

Table 5.2.: Guidelines based on the temperature difference

5.3.4. INFLUENCE OF THE SOLAR RADIATION ON THE INTERNAL SURFACE TEMPERATURE


To evaluate the influence of the sun on the internal surface temperature, WR5E is simulated again with a
constant indoor and outdoor temperature that has a difference of 15 degrees Celsius. Also the surface
heat transfer coefficient is taken as a constant to ignore any influence of the wind. The sun is simulated.
Further explanation is given for each box plot. Next to each explanation of the box plot, a number of
values are given. This is the number of one hour periods that satisfy these criteria.
0,94 0,94 0,93
0,90

0,80
Temperzture factor

0,70 0,67
0,63

0,60

0,55 0,56 0,55 0,55 0,55


0,50

0,40

0,30
A B C D E F
Figure 5.9: Temperature factor distribution depending on guidelines for the solar radiation

A: The dynamic temperature factors of the WR5E junction simulated with the realistic
boundary conditions. This is the same as the WR5E junction in figure 5.4. The steady
state temperature factor is 0.6. (4729 values)
B: The dynamic temperature factors as a result of the numerical calculation as described
above. Only the sun as an influence factor. (4729 values)
C: The dynamic temperature factors only measured when there was no solar radiation
during this simulation. (2838 values)
D: The dynamic temperature factors only measured when there was solar radiation during
this simulation. (1891 values)

63
Chapter 5| Development of guidelines to detect different building envelope interfaces

E: The dynamic temperature factors only measured when the daily absorbed solar energy
on the outside surface, not exceed 1000 w/m 2. If so, then al values calculated in the next
16 hours where ignored. (2536 values)
F: The dynamic temperature factors that suffice the guidelines described in table 5.3. Al
time steps that are not according to this table where neglected. (2706 values)

Daily absorbed solar energy Time that no infrared imaging should be taken
500 W/m2-1000 W/m2 4 hours after last direct solar irradiation
2 2
1000 W/m -1500 W/m 8 hours after last direct solar irradiation
2 2
1500 W/m -2000 W/m 12 hours after last direct solar irradiation
2 2
2000 W/m -2500 W/m 14 hours after last direct solar irradiation
2 2
2500 W/m -3000 W/m 16 hours after last direct solar irradiation
2 2
3000 W/m -3500 W/m 18 hours after last direct solar irradiation
3500 W/m2 and more 20 hours after last direct solar irradiation
Table 5.3: Guidelines based on the solar radiation influence

If only the sun is simulated and the rest of the boundary conditions are set constant, the values between
the 5% fractal and the 50% fractal are less distributed compared to realistic conditions. [box plot A
compared with B,C,D and E]. Guidelines that describe that infrared imaging can only be takien if there is
no direct solar radiation are not correct. This is supported by box plot C and D were there is a negligible
difference between the values that are measured while there was direct solar radiation and when there
was only no direct solar radiation. This means that the period after direct solar radiation has an influence
on the temperature factors on the internal surface of a building envelope interface. Box plot E an F gives
the most accurate results. The most detailed guidelines, these of box plot F, give values between 0.56 and
0.62 for respectively 5% fractal and 95% fractal. This means there is only a deviation of 0.03 compared to
the steady state temperature factor. Realistically, it is not only impractical to measure the daily absorbed
solar energy, it is not feasible. The requirements that are described in table 5.3 are not feasible to use in
practice. These results give us a better view of the impact of solar radiation but can't be used in practice.
Therefore, more applicable guidelines were composed that are more useful for in practice. After further
examination of the results, the main conclusion is that only after a couple of hours, the solar radiation has
an influence on the internal surface temperature. Not only during the direct solar radiation but also after,
the impact has to take into consideration. Therefore, as shown in figure 5.10, the distribution of the

64
Chapter 5| Development of guidelines to detect different building envelope interfaces

temperature factor is examined during the morning because that is the moment the sun comes up and
temperature values shows the smallest deviation. Different time gaps were taken into consideration as
shown in figure 5.10.

1,02
1,00 0,83 0,80
0,80 0,66
0,60
Temperature factor

0,40
0,20
0,29
0,00
-0,20
-0,12 -0,12
-0,40
-0,60
-0,80
-1,00
-0,96
-1,20
Realistic 6-12 am 7-12 am 8-12 am

Figure 5.10: Temperature factor distribution depending on time interval

realistic 6-12am 7-12am 8-12 am


5% fractal-95%fractal 0,41-0,73 0,43-0,69 0,43-0,66 0,43-0,61
Table 5.4: 5% fractal-95%fractal values shown in figure 5.10

The measurements that are performed during the heating season between 8am and 12am are less
accurate then the guidelines described in table 5.3. This method is more feasible and still more accurate
than when measurements were done without guidelines.

5.3.5. INFLUENCE OF AL GUIDELINES ON THE INTERNAL SURFACE TEMPERATURE


Simulations that combine these guidelines for wind speed (no guidelines), temperature difference (10°C,
15°C and 18°C°) and solar radiation (only between 8am and 12 am) are applied on the WR5E junction and
shown in figure 5.11. The reference simulation is the original one that is shown in figure 5.4.

65
Chapter 5| Development of guidelines to detect different building envelope interfaces

0,70
0,66 0,65 0,65 0,65 0,65
0,65

0,60

Temperature factor
0,55

0,50

0,45
0,46 0,47
0,40
0,40
0,35

0,30 0,34

0,25 0,29

0,20
8-12 am 8-12 am, 8-12 am, 8-12 am, 8-12 am,
ΔT>10 ΔT>15 ΔT>18 ΔT>20
Figure 5.11: Temperature factor distribution depending on time interval and ΔT

Realistic ΔT >10°C ΔT >15°C ΔT >18°C ΔT>20


5% fractal-95%fractal 0,41-0,73 0,44-0,61 0,47-0,61 0,50-0,61 0,50-0,61
Table 5.5: 5% fractal-95%fractal values shown in figure 5.11

The values shown in table 5.5 are only the values measured between 8am and 12 am. There are no
differences between the values when the temperature differences were higher than 18°C or higher than
20°C degrees. The most accurate approach to determine the temperature value on the internal surface
with infrared imaging is to do the survey between 8 am and 12 am with a temperature difference of at
least 18°C. Between the 5% fractal and the 95% fractal is only 0.11 difference while if you performing a
random measurement, this 0.32 is. These guidelines are applied on the rest of the wall roof connections.
For the temperature difference, the other two guidelines are also taken into consideration.

5.3.6. GUIDELINES APPLIED ON ALL WALL/ROOF JUNCTIONS


Table 5.6 presents the values between the 5% and 95% fractal of all temperature factors that were
calculated during the period that meets the guidelines. These are the guidelines for the temperature
difference and the solar radiation (between 8am and 12 am). There were no restrictions applied on the
wind speed. These are compared to the steady state temperature factor and the distribution of all
temperature factors calculated during the whole heating season without any restrictions.

66
Chapter 5| Development of guidelines to detect different building envelope interfaces

Normal ΔT>10°C ΔT>15°C ΔT>18°C


f-factor
5%-95% Range 5%-95% Range 5%-95% Range 5%-95% Range
WR0E 0.57 0.38-0.72 0.34 0.41-0.59 0.18 0.44-0.59 0.15 0.46-0.59 0.13
WR0 0.60 0.41-0.73 0.32 0.44-0.61 0.17 0.46-0.61 0.15 0.49-0.61 0.12
WR5E 0.60 0.41-0.73 0.32 0.44-0.62 0.18 0.47-0.62 0.15 0.49-0.62 0.13
WR5 0.68 0.49-0.76 0.27 0.54-0.70 0.16 0.55-0.69 0.14 0.58-0.69 0.11
WR14E 0.74 0.59-0.85 0.26 0.64-0.80 0.16 0.65-0.79 0.14 0.67-0.80 0.13
WR14 0.86 0.74-0.95 0.21 0.79-0.93 0.14 0.79-0.91 0.12 0.81-0.91 0.10
WR14NTB 0.91 0.79-1.00 0.21 0.85-0.99 0.14 0.85-0.97 0.12 0.87-0.97 0.10
WR24 0.90 0.78-0.99 0.21 0.83-0.97 0.14 0.84-0.96 0.12 0.85-0.95 0.10
WR24NTB 0.94 0.83-1.04 0.21 0.88-1.02 0.14 0.88-1.00 0.12 0.90-1.00 0.10
Table 5.6:Applied guidelines on all wall/roof junctions: steady state f-factor, 5% fractal-95%fractal values and range are presented

The most stringent guidelines where the temperature difference is at least 18°C between 8am and 12 am
is shown in figure 5.12.
1,05 1,02
0,99 0,98
1,00
0,94
0,95

0,90

0,85 0,82
0,87
0,80 0,84 0,82
Temperature factor

0,75 0,72 0,78


0,70 0,65
0,65
0,63
0,65

0,60 0,63
0,55

0,50 0,54

0,45
0,46 0,46
0,40 0,43
0,35
WR0E WR0 WR5E WR5 WR14E WR14 WR14NTB WR24 WR24NTB
Figure 5.12: Applied guidelines for the sun and ΔT>18°C on all wall/roof junctions

Even with the most stringent requirements, not all building envelope interfaces can be separated. There
is almost no difference between the results of the WR0 and the WR5E junction. There is more chance to
get different results between WR and WR5. The most clear difference is between the WR0, WR14E and

67
Chapter 5| Development of guidelines to detect different building envelope interfaces

WR14 junction. If these results are combined with the outside results, the most of the building envelope
interfaces can be distinguished. These results are shown in figure 5.3. For example differentiate the
temperature factors in the corners of the WR0 and WR5E junctions is not possible with infrared imaging.
If the temperature factor values are between the 95% fractal value of the WR0 junction and the 5%
fractal value of the WR14 junction, than can be concluded that the building envelope interface is a
junction with 5 centimeters of insulation. Only if the measurements outside detect a large difference
between the temperature factor in point 6 and point 8, the it is an temperature value of the WR14E
junction

As mentioned before, outside temperature factors can be used to distinguish the difference between an
extended concrete roof slab or a normal concrete slab. The guidelines for an infrared thermograph survey
performed on the external surface of a building envelope are presented in the next paragraph about
balconies.

Conclusion
Te determination of the different wall/roof connections can be divided in two groups, one with high
thermal performance and a second with lower thermal performance. Further detailed analyses of the
junctions is possible but an infrared thermography survey on the external surface has to be performed to
distinguish a junction with or with an extended concrete roof slab. Infrared thermography should not be
used to distinguish the wall roof junctions with 5 and 0 centimeter of insulation, because there is a great
chance that these values overlap, even with the stringent guidelines.

5.4. Balcony connection

5.4.1. MEASUREMENTS ON THE INTERNAL SURFACE


The guidelines that are represented in paragraph 5.3 are also applied on the balcony connection. Figure
5.13 shows the distribution of all temperature factors during the heating season with realistic boundary
conditions. Six different balcony connections are presented in the chart. As in previous paragraphs, the
names of the building envelope interfaces are displayed at the x-axes.

68
Chapter 5| Development of guidelines to detect different building envelope interfaces

1,35 1,38
1,40

1,20 1,11
1,03
1,00 0,91
0,78
0,80
Temperature factor

0,60
0,59 0,63
0,40

0,20

0,00

-0,20 -0,08

-0,40 -0,25

-0,60 -0,47
B0 B5 B14 B14NTB B24 B24NTB

Figure 5.13: Temperature factor distribution in point 3 of the six balcony junctions

The distributions of the temperature factors are more separated in comparison to results of the wall/roof
junctions without applying any guidelines. Some box plots do overlap, for example, a value measured for
the B5 connections can be interpreted as a value measured on the internal surface of a B24NTB
connection. Therefore, the guidelines for temperature difference (ΔT>10°C, ΔT>15°C, ΔT>18°C) and solar
radiation (between 8-12 am) on the internal surface temperature factor in point three are applied and
illustrated in Figure 5.14-16. As mentioned before, there are no guidelines for wind speed necessary for
indoor measurements.

1,20 1,12 1,14 1,20 1,08 1,10 1,20


1,02 1,04
0,99 0,97
0,94 1,00 0,92 1,00 0,91
1,00 0,87
0,85 0,84
0,80
0,80 0,80 0,80 0,88
0,83 0,86
0,62 0,81 0,83 0,82
0,59 0,76
0,56 0,72
0,60 0,71 0,60 0,71 0,60
0,66 0,68
0,61 0,64
0,56
0,40 0,40 0,40

0,20 0,20 0,20 0,33


0,29

0,14
0,00 0,00 0,00

Figure 5.14: ΔT>10°C between 8-12am Figure 5.15: ΔT>15°C between 8-12am Figure 5.16: ΔT>18°C between 8-12am

69
Chapter 5| Development of guidelines to detect different building envelope interfaces

Table 5.7 presents the calculated values between the 5% and 95% fractal of all temperature factors
shown in Figure 5.14-5.16. The listed range is the difference between 5% and 95% fractal. These results
indicate that no overlap is observed between the values of certain balcony connections when the
guidelines are used. There is an explicit difference between values measured for a balcony with or
without a thermal bridge and this was not the case before application of the guidelines. A clear difference
between B0, B5 and B24 connections can be observed while the contrast between the B5 and B14 and
the B14 and B 24 junctions is not always obvious.

Normal ΔT>10°C ΔT>15°C ΔT>18°C


5%-95% Range 5%-95% Range 5%-95% Range 5%-95% Range
B0 0.20-0.6 0.4 0.29-0.54 0.25 0.33-0.54 0.21 0.37-0.54 0.17
B5 0.59-0.83 0.24 0.63-0.79 0.16 0.65-0.78 0.13 0.67-0.78 0.11
B14 0.66-0.89 0.23 0.71-0.86 0.15 0.73-0.85 0.12 0.74-0.85 0.11
B14NTB 0.83-1.04 0.21 0.87-1.02 0.15 0.88-1.00 0.12 0.89-1.00 0.11
B24 0.7-0.93 0.23 0.76-0.91 0.15 0.77-0.89 0.12 0.78-0.89 0.11
B24NTB 0.84-1.05 0.21 0.89-1.04 0.15 0.90-1.02 0.12 0.91-1.02 0.11
Table 5.7: Applied guidelines on internal of all balcony connections: 5% fractal-95%fractal values and range are presented

5.4.2. MEASUREMENTS ON THE EXTERNAL SURFACE


In the next paragraph, the influence of boundary conditions on the external surface is discussed.
Therefore, the balcony connections are simulated with computer software Voltra. One of the connections
is further investigated in order to test guidelines that where proposed in chapter four and to develop new
recommendations. The temperatures are calculated in points 7 and 10 of the balcony building envelope
interface. These points are defined in figure 5.1. The temperature factor is used as a dimensionless value
to compare the influences of wind speed, solar radiation and temperature difference between inside and
outside.

70
Chapter 5| Development of guidelines to detect different building envelope interfaces

0,60 0,60

0,50 0,50

0,40 0,40

0,30 0,30
Temperature factor

Temperature factor
0,20 0,20

0,10 0,10

0,00 0,00

-0,10 -0,10
-0,07
-0,20 -0,20 -0,14 -0,12
-0,15
-0,30 -0,30 -0,24 -0,24
-0,40 -0,40
B0 B5 B14 B14NTB B24 B24NTB B0 B5 B14 B14NTB B24 B24NTB
Figure 5.17: f-factor distribution in point 7 Figure 5.18: f-factor distribution in point 10

Figure 5.17 and 5.18 show the f-factor distribution of the values calculated in point 7 and 10. Point 10 is
located on top of the balcony while point 7 is at the bottom where solar radiation should have less
impact. This can explain the difference between the two graphs. The next paragraph will investigate this
phenomenon. To determine new guidelines for infrared thermography imaging, the B5 connections are
used for further simulations. Different boundary conditions will be investigated by several simulations on
the B5 connection. All boundary conditions were kept constant during this experiment, except for one. By
varying a certain boundary condition, its influence on the temperature factor will become visible in the
results. The simulation of each boundary condition is performed in point 7 and in point 10, so the impact
of an overhang of the balcony can be evaluated by comparing both results. The outcome is represented in
figure 5.19 and table 5.8.

1,00
0,90
0,80 0,68
0,70
0,60
Temperature factor

0,50 0,37 0,37


0,40
0,30 0,19 0,19
0,20
0,10
0,14 0,14
0,00 0,07 0,07
-0,10
-0,20
-0,30
-0,40
A B C D E F
Figure 5.19: f-factor distribution calculated at point 7 and point 10 with all boundary conditions constant except for one that is
fluctuating

71
Chapter 5| Development of guidelines to detect different building envelope interfaces

A and B: Results at respectively point 7 and point 10 for fluctuating indoor and outdoor
temperatures. No solar radiation and variable wind speed is taken into account.
C and D: Results at respectively point 7 and point 10 where only the solar radiation is simulated.
Wind speed and the internal and external temperatures are constant.
E and F: Results at respectively point 7 and point 10 for fluctuating wind speed with no impact of
solar radiation or temperature fluctuations.
A B C D E F
Under Above Under Above Under Above
5%-95% fractal -0,20-0,17 -0,20-0,17 0,15-0,38 0,15-0,65 0,09-0,17 0,09-0,17
Table 5.8: 5%-95% fractal of the f-factor distribution calculated a point 7 and point 10 as shown figure 5.19

The results indicate that the overhang of the balcony has no influence on the wind speed or temperature
difference. Furthermore, a different distribution of the f factor is observed when the calculations are
done under or on top of the balcony. Therefore further simulations will be done in point 10 because this
is most influenced by the solar radiation. The impact of an overhang on the surface temperature is out
the scope of this thesis. The wind speed influences the f-factor much less than the solar radiation or the
temperature difference between inside and outside. These findings are further investigated in the next
paragraphs.

5.4.3. INFLUENCE OF THE WIND SPEED ON THE EXTERNAL TEMPERATURE FACTOR


The graphic in figure 5.20 shows the values of the temperature factor depending on the wind speed. Solar
radiation was not taken into consideration during these simulations. The temperature difference between
indoor and outdoor was also kept constant. If the wind speed increases, the temperature factor will
decrease as they are inversely proportional.
0,20

0,18

0,16
Temperature factor

0,14

0,12

0,10

0,08

0,06
0 2 4 6 8 10 12 14
Figure 5.20: f-factor distribution calculated at point 10 with all boundary conditions constant except for the wind speed

72
Chapter 5| Development of guidelines to detect different building envelope interfaces

5.4.4. INFLUENCE OF THE TEMPERATURE DIFFERENCE BETWEEN INDOOR AND OUTDOOR ENVIRONMENT ON THE
EXTERNAL TEMPERATURE FACTOR.

The distribution of the f-factor depending on the temperature difference between indoor and outdoor is
represented in figure 5.21. This is similar to the chart where the temperature factor was calculated for
the internal surface . The same guidelines are applied for outdoor measurements.

0,50

0,00

-0,50
Temperature factor

-1,00

-1,50

-2,00
0 5 10 15 20 25 30 35

-2,50

Figure 5.21: f-factor distribution calculated at point 10 with all boundary conditions constant except for the temperature

5.4.5. INFLUENCE OF THE SOLAR RADIATION ON THE EXTERNAL TEMPERATURE FACTOR


1,30 1,23 1,23 1,23

1,20
1,06
1,10
1,01
1,00

0,90
Temperature factor

0,80

0,70
0,59
0,60
0,49
0,50

0,40
0,30
0,30 0,25

0,20

0,10 0,15 0,15 0,15


0,14 0,14 0,14 0,14 0,14 0,15
0,00
A B C 0-4am 4-8am 8-12am 0-4pm 4-8pm 8-12pm
Figure 5.22: f-factor distribution calculated at point 10 with all boundary conditions constant except for the solar radiation

73
Chapter 5| Development of guidelines to detect different building envelope interfaces

0-4am 4-8am 8-12am 0-4pm 4-8pm 8-12pm


5% fractal - 95% fractal 0,15-0,24 0,15-0,20 0,15-0,60 0,17-0,94 0,16-0,67 0,15-0,34
Range 0,09 0,05 0,45 0,77 0,51 0,19
Table 5.9: 5%-95% fractal of the f-factor distribution calculated at point 10 as shown in the figure 5.22

A: The dynamic temperature factors of the B5 connection, simulated with only solar radiation.
Internal and external temperatures and the wind speed were set constant.
B: The dynamic temperature factors of the B5 connection, simulated with only solar radiation.
Internal and external temperatures and the wind speed were set constant. Only the values
where no direct solar radiation was measured are shown.
C: The dynamic temperature factors of the B5 connection, simulated with only solar radiation.
Internal and external temperatures and the wind speed were set constant. Only the values
where direct solar radiation was measured are shown.

The examination of solar radiation on the external surface is shown in figure 5.22. There is a clear
difference between the values calculated with or without direct solar radiation. This conclusion is
opposite to the result of the internal surface temperature, where no difference is seen. This dissimilarity
can be explained by means of the thermal inertia of the construction because it has more influence on
the internal surface temperature. Furthermore, the temperature factors were divided in different time
intervals so these guidelines can be easily used in practice. The diagram shows that the best time to do a
thermographic survey on the external surface is between 4 and 8 o’clock in the morning, just before
sunrise. In the next part of this paper, different guidelines are simultaneously applied on the B5
connection.

74
Chapter 5| Development of guidelines to detect different building envelope interfaces

5.4.6. INFLUENCE OF AL GUIDELINES ON THE EXTERNAL SURFACE TEMPERATURE


1,00

0,90

0,80

0,70

0,60
Temperature factor

0,50

0,40
0,30 0,30
0,30 0,24
0,20
0,16
0,20

0,10

0,00
0,04 0,04
-0,10 -0,02 -0,02 -0,01
-0,07
-0,20
Realistic 4-8 am 4-8am, ΔT>10°C 4-8am, ΔT>15°C 4-8am, ΔT>18°C 4-8am, ΔT>20°C
Figure 5.23: f-factor distribution where different guidelines are applied on the B5 connection

4-8am, 4-8am, 4-8am, 4-8am,


Realistic 4-8 am ΔT>10°C ΔT>15°C ΔT>18°C ΔT>20°C
5% fractal - 95% fractal 0,05-0,48 0,03-0,2 0,03-0,19 0,04-0,18 0,07-0,16 0,07-0,15
Range 0,43 0,17 0,16 0,14 0,09 0,08
Table 5.10: 5%-95% fractal of the f-factor distribution calculated at point 10 as shown in the figure 5.23

Figure 5.23 and table 5.10 show the results of the applied guidelines for solar radiation (values between 4
and 8 am) and temperature difference between indoor and outdoor. Guidelines for the sin speeds are not
applied because there is almost no change in distribution of the temperature factor. The most stringent
guidelines, these of 4-8am and temperature difference not lower than 18°C, are applied on al balcony
connections.

5.4.7. GUIDELINES ON AL BALCONY CONNECTION


Even with very strict guidelines, the different balcony connections are not always visible on an infrared
thermography image that is taken on the external surface. Results are shown for the values in point 7 and
point 10. Even after applying the guidelines(measurements are done between 4-8 am and the
temperature differential cannot be lower than 18°C), there is no difference in distribution of temperature
factor.

75
Chapter 5| Development of guidelines to detect different building envelope interfaces

0,25

0,20
0,20 0,18
0,17
0,16

0,15
0,12
0,10
0,10
Temperature factor

0,05

0,04
0,00 0,02 0,02
0,00

-0,05

-0,10 -0,08
-0,09

-0,15
B0 B5 B14 B14NTB B24 B24NTB

Figure 5.24: f-factor distribution of al balcony connections with the applied guidelines in point 7

0,25
0,20
0,20 0,18
0,17
0,16

0,15 0,12
0,10
Temperature factor

0,10

0,05

0,04
0,00 0,02
0,02
0,00
-0,05

-0,10 -0,08
-0,09

-0,15
B0 B5 B14 B14NTB B24 B24NTB
Figure 5.25: f-factor distribution of al balcony connections with the applied guidelines in point 10

76
Chapter 6

Conclusion and recommendations

6.1. General conclusion


This thesis presents the comparison between steady state and transient calculations of heat losses
through the building envelope interfaces in order verify the use of the linear thermal transmittance
factor. This is an answer to the ongoing debate in North America. Thermal bridges and 2 dimension heat
losses are not standard included in their calculations and a lot of research is going on about this topic. The
total heat loss trough different building envelopes was compared between steady state and transient
conditions. This was done in order to show the reliability of the steady state approach for transient
conditions by using the degree hour method. The difference between the total calculated heat losses
when using steady state and transient approaches is due to the thermal inertia of the materials. Although
there is a little difference between the results in steady state and transient conditions, it was found to be
negligible. The comparison of the heat loss trough the junctions shows us that the linear thermal
transmittance calculated under steady state conditions is a very good approximation of the two
dimensional heat losses through a building envelope in transient conditions. This is a useful calculation
method that is more feasible than the complicated numerical calculation methods that are typically used
in North America.

The second part the thesis presents guidelines for the boundary conditions when detecting thermal
bridges through the use of infrared thermography images. The guidelines provide restrictions on the wind

77
Chapter 6| Conclusion and recommendations

speed, solar radiation and temperature difference between indoor and outdoor to ensure a more
accurate approach of the temperature values. By using these restrictions, the temperature factor can be
measured only during certain periods of the day..

A first remark on the temperature factor is that the average of the f-factor in dynamic conditions is lower
than when calculated in steady state. The difference depends on the thermal performance of the
connections or junctions, solar radiation and surface heat transfer coefficient. The better the thermal
performance or the smaller the thermal bridge is, the smaller this difference. This may be due to the fact
that the structure is less influenced by the fluctuating outdoor conditions because of the insulation layer.

When an infrared thermal survey is completed without guidelines, there is no difference between the
different thermal performance building envelope interfaces. The only distinction that can be made is
between building envelope interfaces that satisfy the EPB regulations and building envelope intersection
that have a lower thermal performance.

The rules that are provided separate the indoor and outdoor use of an infrared camera. In general, wind
speed has almost no influence on the result and therefore no restrictions are presented. A second
boundary conditions that is evaluated is the temperature difference between inside and outside. For
images taken outside as for images taken of the internal surface, the temperature differential cannot be
lower than 10 °C. This is a very optimistic value and a recommended value of 15 °C should be used while a
more stringent differential temperature of at least 18°C gives the best results. With respect to solar
radiation, practical recommendations on inside infrared thermography are that imaging can only be
conducted between 8 and 12 am and on the outside this is between 4 and 8 am.

After further analyzing the results, the guidelines can give a rough estimation of the temperature factor
but infrared is not accurate enough to detect different building envelope interfaces. Only an estimation
of the thermal performance of the junction can be made. This is most evident on the external surface,
where the difference between temperature factors is not obvious enough. For indoor measurements, a
good estimation of the temperature value is made and the difference between the different junctions can
be noticed. Although the overlap between the same f-factors can be measured while applying the
guidelines, the f-factors that are measured for a wall/roof junction with no insulations can still be a wall
roof junction with 5 centimeter of insulations.

78
Chapter 6| Conclusion and recommendations

Thermal bridges in balcony and wall/roof junctions can be detected by using infrared thermography but
only if this applied on the internal surface of the building envelope. Infrared imaging applied on the
external surface can detect if the concrete slab is in contact with the outer leaf cavity wall or not. Infrared
thermography can also give a rough estimating of the temperature factor and can be used to assess the
risk for growth mould, but with an error of maximum 0.07 for wall/roof junctions and 0.09 for balcony
connections. These errors are based on the building envelope interface where no insulations are used.

6.2. Recommendations
In chapter five, a first step is given for a more detailed approach for the impact of solar radiation on the
building envelope. But this is not feasible for practical use. More examination of solar radiation can
provide more detailed and useful guidelines that can be applied. Also the influence of overhangs can be
further investigated even as the different orientation. Further research to develop tables with possible
errors for different building envelope interfaces so reliable estimation can be done. This would be a great
asset planning building retrofits.

79
Chapter 6| Conclusion and recommendations

80
Bibliography

Bibliography

[1] A. M. Omer, “Energy, environment and sustainable development,” Renew. Sustain. Energy Rev., vol. 12, no.
9, pp. 2265–2300, Dec. 2008.

[2] H. Ge, V. R. McClung, and S. Zhang, “Impact of balcony thermal bridges on the overall thermal performance
of multi-unit residential buildings: A case study,” Energy Build., vol. 60, pp. 163–173, May 2013.

[3] EN ISO 10211, “Thermal bridges in building construction - Heatflows and surface temperatures - Detailed
calculations,” 2008.

[4] B. Lehmann, K. Ghazi Wakili, T. Frank, B. Vera Collado, and C. Tanner, “Effects of individual climatic
parameters on the infrared thermography of buildings,” Appl. Energy, vol. 110, pp. 29–43, Oct. 2013.

[5] C. a. Balaras and a. a. Argiriou, “Infrared thermography for building diagnostics,” Energy Build., vol. 34, no.
2, pp. 171–183, Feb. 2002.

[6] Physibel, VOLTRA v7.0w Computer software. 2011.

[7] H. Hens, Building Physics - Heat, Air and Moisture: Fundamentals and Engineering Methods with
Examplesand Exercises. 2012.

[8] Y. Gao, J. J. Roux, L. H. Zhao, and Y. Jiang, “Dynamical building simulation: A low order model for thermal
bridges losses,” Energy Build., vol. 40, no. 12, pp. 2236–2243, Jan. 2008.

[9] K. Martin, C. Escudero, a. Erkoreka, I. Flores, and J. M. Sala, “Equivalent wall method for dynamic
characterisation of thermal bridges,” Energy Build., vol. 55, pp. 704–714, Dec. 2012.

[10] A. Janssens, E. Van Londersele, B. Vandermarcke, S. Roels, P. Standaert, and P. Wouters, “Development of
Limits for the Linear Thermal Transmittance of Thermal Bridges in Buildings,” Therm. Perform. Exter. Envel.
Whole Build. X Int. Conf., 2007.

[11] EN ISO 14683, “Thermal bridges inbuilding construction - Linear thermal transmittance - Simplified
methods and defaults values,” 2008.

[12] J. A. Duffie and W. A. Beckman, Solar Engineering of Thermal Processes, Third Edit. 2006.

[13] A. H. Taki and D. L. Loveday, “External convection coefficients for framed rectangular elements on building
facades,” Energy Build., vol. 24, no. 2, pp. 147–154, Jul. 1996.

[14] D. L. Loveday and A. H. Taki, “Convective heat transfer coefficients at a plane surface on a full-scale
building facade,” Int. J. Heat Mass Transf., vol. 39, no. 8, pp. 1729–1742, May 1996.

[15] EN ISO 6946, “Building components and building element - thermal resistance and thermal transmittance -
Calculation method,” 2008.

[16] A. Janssens, Bouwfysica: Bouwfysische aspecten van gebouwen. 2010.

81
Bibliography

[17] ISO/FDIS 13786, “Thermal performance of building components — Dynamic thermal characteristics —
Calculation methods,” vol. 2006, no. 50, 2006.

[18] M. Ozel, “Effect of insulation location on dynamic heat-transfer characteristics of building external walls
and optimization of insulation thickness,” Energy Build., vol. 72, pp. 288–295, Apr. 2014.

[19] S. A. Al-Sanea and M. F. Zedan, “Effects of insulation location on thermal performance of building walls
under steady state periodic conditions,” Int. journey Ambient energy 22, 2001.

[20] S. A. Al-Sanea, “Effect of insulation location on initial transient ther-mal response of building walls,” Int.
journey Ambient energy 24, 2001.

[21] S. M. Zubair and E. Al-Regib, “Transient heat transfer through insulated walls,” Energy 20, 1995.

[22] K. J. Kontoleon and E. A. Eumorfopoulou, “The influence of wall orientation and exte-rior surface solar
absorptivity on time lag and decrement factor in the Greekregion,” Renew. enrgy 33, 2008.

[23] D. K. Bikas and K. J. Kontoleon, “The impact of temperature variences on thermal iner-tia factors of opaque
elements of the building envelope,” Eight Int. Conf. Environ. Sci. Technol. Lemn. Isl., 2003.

[24] F. Aguilar, J. P. Solano, and P. G. Vicente, “Transient modeling of high-inertial thermal bridges in buildings
using the equivalent thermal wall method,” Appl. Therm. Eng., vol. 67, no. 1–2, pp. 370–377, Jun. 2014.

[25] K. Sedlbauer, Vorhersage van schimmelpilzbildung auf und un bauteilen. 2001.

[26] H. Hens, Fungal defacement in buildings: A performance related approach. 1999.

[27] O. Adnan, On the fungal defacement of interior finishes. 1994.

[28] P. Johansson, A. Ekstrand-Tobin, T. Svensson, and G. Bok, “Laboratory study to determine the critical
moisture level for mould growth on building materials,” Int. Biodeterior. Biodegradation, vol. 73, pp. 23–
32, Sep. 2012.

[29] “Maximaal toelaatbare U waarden of minimaal te realiseren R waarden,” in Vlaams Energieagentschap,


2014.

[30] Physibel, VOLTRA v7.0w Manual. 2011, pp. 1–55.

[31] Physibel, TRISCO and KOBRU v12.0w Manual, vol. 25, no. 1. 2010.

[32] B. T. Griffith, D. Türler, and D. Arasteh, “Surface Temperatures of Insulated Glazing Units : Infrared
Thermography Laboratory Measurements,” no. December 1995, 1996.

[33] Griffith, B. T. . Beck, D. Arasteh, and D. Turler, “Issues associated with the use of infrared thermography for
experimental testing of insulated systems,” 1995.

[34] ASTME1862-97, “Standard Test Method for Measuring and Compansating for reflected Temperaturure
Using Inrared Imaging Radiometers,” Astm Am. Soc. Test. Mater., 2002.

[35] B. C. Pearson, “Thermal Imaging of Building Fabric References,” 2011.

82
Bibliography

[36] ASTME1897-97, “Standard Test Method for Measuring and Compansating for Transmittance of an
Attenuating Medium Using Infrared Imaging Radiometers,” Astm Am. Soc. Test. Mater., 2002.

[37] P. A. Fokaides and S. A. Kalogirou, “Application of infrared thermography for the determination of the
overall heat transfer coeffcient in building envelopes,” Appl. Energy, vol. 88, pp. 4358–4365, 2011.

[38] R. Albatici and A. Tonelli, “Infrared thermovision technique for the assesment of thermal transmitance
value of opaque buidling elements on site,” Energy Build., vol. 42, pp. 2177–2183, 2010.

83
84
List of tables

Table 2.1: Conventional surface resistance 14


Table 2.2: Maximum thermal transmittance values 23
Table 3.1: Designation of the building envelope interface 30
Table 3.2: Comparison of the results of the building envelope interface WR14 41
Table 3.3: Total heat loss wall/roof junction and the percentage through the junction in steady 42
state and dynamic
Table 3.4: Total heat loss balcony connection and the percentage through the connection in 42
steady state and dynamic
Table 3.5: Total heat loss wall/floor junction and the percentage through the junction in steady 43
state and dynamic
Table 4.1: limitations for the wind speed 50
Table 4.2: Limitations for the temperature difference 50
Table 4.3: Limitations for the solar radiation 51
Table 5.1: f-Factor of wall/roof junction in steady state and dynamic conditions 57
Table 5.2: Guidelines based on the temperature difference 63
Table 5.3: Guidelines based on the solar radiation influence 64
Table 5.4: 5% fractal-95%fractal values shown in figure 5.10 65
Table 5.5: 5% fractal-95%fractal values shown in figure 5.11 66
Table 5.6:Applied guidelines on all wall/roof junctions: steady state f-factor, 5% fractal- 67
95%fractal values and range are presented
Table 5.7: Applied guidelines on internal of all balcony connections: 5% fractal-95%fractal values 70
and range are presented
Table 5.8: 5%-95% fractal of the f-factor distribution calculated a point 7 and point 10 as shown 72
figure 5.19
Table 5.9: 5%-95% fractal of the f-factor distribution calculated at point 10 as shown in the 74
figure 5.22
Table 5.10: 5%-95% fractal of the f-factor distribution calculated at point 10 as shown in the 75
figure 5.23

85
List of figures

Figure 2.1: Building section with conventions for the one-dimensional geometrical reference. 10
Figure 2.2: Building section between wall and roof with a point thermal bridge in the corner. 10
Figure 2.3: Non-periodically boundary conditions 14
Figure 2.4: Periodically boundary conditions 14
Figure 2.5: Penetration depth 16
Figure 2.6: Equivalent wall method 19
Figure 3.1: Colours window Voltra 26
Figure 3.2: Grid Auto split 27
Figure 3.3: Linear grid of 25mm meshes 28
Figure 3.4: Auto split used a grid 28
Figure 3.5: Text output 29
Figure 3.6: Graphical output of the surface temperature distribution 30
Figure 3.7: Example of a building envelope interface 31
Figure 3.8 and 3.9: Wall/roof junctions with no insulation 32
Figure 3.10 and 3.11: Wall/roof junctions with 5cm retrofit cavity insulation 32
Figure 3.12, 3.13 and 3.14: Wall/roof junctions with 14 cm insulation 32
Figure 3.15 and 3.16: Wall/roof junctions with 24 cm insulation 33
Figure 3.17 and 3.18: Balcony connection with 0 and 5 cm insulation 33
Figure 3.19 and 3.20: Balcony connection with 14 cm insulation 34
Figure 3.21 and 3.22: Balcony connection with 24 cm insulation 34
Figure 3.23 and 3.24: Wall/floor junction with 0 cm insulation 35
Figure 3.25 and 3.26: Wall/floor junction with 5 cm insulation 35
Figure 3.27 and 3.28: Wall/floor junction with 14 and 24 cm insulation 36
Figure3.29: Variable in and outdoor temperature 37
Figure3.30: Variable outdoor climate date for the heating season, Uccle, Belgium and the variable 37
indoor temperate regimes measured in a dwelling of a Belgian family
Figure3.31: WR14 building envelope interface 39
Figure 3.32: Heat flow in and out of the object 40
Figure 3.33: Heat flow in and out of the building envelope interface 41

86
Figure 4.1: Schematization of the heat transfer process between infrared camera and the object. 46
Figure 4.2: IR picture with scale between 13.8°C and 17.5°C 48
Figure 4.3: IR picture with more detailed scale between 17°C and 17.7°C 48
Figure 5.1 Balcony connection with surface temperatures 54
Figure 5.2 Wall/roof connection with surface temperatures 54
Figure 5.3 Box plot distribution 56
Figure 5.4: f-Factor distribution of the different wall/roof junctions in point 3 58
Figure 5.5: F1 minus F3 distribution of the different wall/roof junctions 59
Figure 5.6: WR5E f-factor distribution different parameters 60
Figure 5.7: All dynamic temperature factors against the temperature difference 61
Figure 5.8: Temperature factor distribution within 1⁰C temperature difference bin 62
Figure 5.9: Temperature factor distribution depending on guidelines for the solar radiation 63
Figure 5.10: Temperature factor distribution depending on time interval 65
Figure 5.11: Temperature factor distribution depending on time interval and ΔT 66
Figure 5.12: Applied guidelines for the sun and ΔT>18°C on all wall/roof junctions 67
Figure 5.13: Temperature factor distribution in point 3 of the six balcony junctions 69
Figure 5.14: ΔT>10°C between 8-12am 69
Figure 5.15: ΔT>10°C between 8-12am 69
Figure 5.16: ΔT>10°C between 8-12am 69
Figure 5.17: f-factor distribution in point 7 71
Figure 5.18: f-factor distribution in point 10 71
Figure 5.19: f-factor distribution calculated at point 7 and point 10 with all boundary conditions 71
constant except for one that is fluctuating
Figure 5.20: f-factor distribution calculated at point 10 with all boundary conditions constant 72
except for the wind speed
Figure 5.21: f-factor distribution calculated at point 10 with all boundary conditions constant 73
except for the temperature
Figure 5.22: f-factor distribution calculated at point 10 with all boundary conditions constant 73
except for the solar radiation
Figure 5.23: f-factor distribution where different guidelines are applied on the B5 connection 75
Figure 5.24: f-factor distribution of al balcony connections with the applied guidelines in point 7 76
Figure 5.25: f-factor distribution of al balcony connections with the applied guidelines in point 10 76

87
88

Anda mungkin juga menyukai