You are on page 1of 227

Building Physics

EXTERIOR CLIMATE
THERMAL COMFORT
HEAT
AIR EXCHANGE
ENERGY
DAYLIGHT
Heinrich Manz
Translated by Carolyn Gorczyca Thrlimann

Lucerne University of Applied Sciences and Arts


School of Engineering and Architecture

Author
Prof. Heinrich Manz, Ph.D.
Lucerne University of Applied Sciences and Arts
School of Engineering and Architecture
Technikumstrasse 21
CH-6048 Horw
Tel. +41 41 349 3915
heinrich.manz@hslu.ch
1. edition, 2014

Sustainable development means development


that meets the needs of the present without
compromising the ability of future generations
to meet their own needs.
Brundtland Commission, World Commission on
Environment and Development (WCED) of the
United Nations (UN), 1987

Content
INTRODUCTION

12

HEAT TRANSFER

14

1.1
1.1.1
1.1.2
1.1.3
1.1.4

Conduction
Heat Equation (Fouriers Differential Equation)
One-Dimensional, Steady-State Heat Conduction
Multi-Dimensional, Steady-State Heat Conduction
One-Dimensional, Non-Steady-State Heat Conduction

14
14
16
17
19

1.2
1.2.1
1.2.2
1.2.3

Convection
Forced and Free Convection
Convection on Surfaces
Convection in Cavities

19
19
20
21

1.3
Thermal Radiation
1.3.1 Laws and Characteristics
1.3.2 Radiation Exchange between Surfaces

23
23
27

1.4

29

Design Values for the Total Heat Transfer on Surfaces

Literature

30

32

EXTERIOR CLIMATE

2.1
2.1.1
2.1.2
2.1.3
2.1.4

Solar Radiation
Solar Radiation Source
Influence of the Atmosphere
Direct Radiation on Tilted Surfaces
Global Irradiance

32
32
33
35
39

2.2

Air Temperatures

43

2.3

Soil Temperatures

45

2.4

Wind

45

2.5

Climate Fluctuations and Changes

46

2.6
2.6.1
2.6.2
2.6.3

Impacts on the Building Envelope


External Heat Transfer Coefficient
Surface Temperature Reduction Due to Infrared Radiation
Service Life of Exterior Building Elements

47
47
47
48

Problems

49

Literature

49

52

THERMAL COMFORT

3.1

Human and Interior Space

52

3.2

Heat Balance of Humans

53

3.3

Requirements for Thermal Comfort

56

3.4
Detailed Comfort Analysis
3.4.1 Global Comfort Criteria
3.4.2 Local Comfort Criteria

58
58
60

3.5

Cold downdraft

62

3.6

Comfort Measurements

64

3.7

Adaptive Comfort

64

3.8

Implications

65

Problems

65

Literature

66

STEADY-STATE THERMAL TRANSMISSION

70

4.1

Standard Cross Section

70

4.2

Thermal Bridges

75

4.3

Cavities in Window Frames etc.

79

Problems

80

Literature

87

NON-STEADY-STATE THERMAL TRANSMISSION

90

5.1
5.1.1
5.1.2
5.1.3

Non-Steady-State Heat Propagation in a Material Layer


Heat Equation and Illustration
Thermal Diffusivity and Thermal Effusivity
Inertia of Walls: Charge and Discharge Performance

90
90
91
93

5.2
5.2.1
5.2.2
5.2.3

Reaction of a Material Layer to Periodic Changes


Material Layer with a Finite Thickness: Amplitude Damping und Phase Shift
Semi-Infinite Material Layers: Penetration Depth
Effective Thickness for Heat Storage

95
95
96
99

5.3

Non-Steady-State Properties of Opaque External Walls

100

5.4

Structural Consequences

104

Problems

105

Literature

106

TRANSPARENT BUILDING ELEMENTS

108

6.1

Classification and Characteristics

108

6.2

Optical Properties of Glazings

110

6.3

Thermal Properties of Glazings

114

6.4

Energy Fluxes Through Windows

119

6.5

Solar Shading Devices

119

Problems

122

Literature

123

AIR EXCHANGE

126

7.1

Wind Pressure on the Building Surface

127

7.2

Thermally Induced Pressure Differences

131

7.3

Airflow Through Leakages

133

7.4

Indoor Air Quality

136

7.5

Airtightness of the Building Envelope

137

7.6

Mechanical Ventilation with Heat Recovery

139

7.7

Passive Cooling by Night-time Ventilation

141

Problems

142

Literature

144

NON-STEADY-STATE BEHAVIOR OF A ROOM

146

8.1

Energy flows in a room

146

8.2

Energy Balance in a Room

148

8.3

Time Constant and Gain/Loss-Ratio

148

8.4

Building Simulation

152

8.5

Building Simulation Example: Office Room in Summer

153

8.6

Structural Consequences

156

Problems

156

Literature

157

ENERGY AND SUSTAINABILITY

160

9.1

Energy and Sustainability Challenges

160

9.2
Heating Energy Demand
9.2.1 Balancing the Energy Flow in a Building
9.2.2 Heating Degree Days

165
165
167

9.3

Protection Against Overheating in Summer

168

9.4

Renewable Energy

169

9.5

Total Energy Expenditure During a Life Cycle

172

9.6

Energy Demand per Floor Area and Energy Standards

174

9.7

Strategies for Low-Energy Buildings

178

9.8

Existing Building Stock and Refurbishment

179

9.9

Climate Change and Energy Demand

180

9.10

Summary

181

Problems

182

Literature

182

10

DAYLIGHT

186

10.1

Solar Radiation and Spectral Luminous Efficiency of the Human Eye

186

10.2
10.2.1
10.2.2
10.2.3
10.2.4
10.2.5
10.2.6

Fundamental Photometric Terms and Relationships


Luminous Flux
Illuminance E
Luminous Intensity I
Photometric Inverse-Square Law
Luminance L
Overview

187
187
189
190
191
192
192

10.3

Luminance Distribution of the Sky

193

10.4

Transmittance Factors of the Building Envelope

194

10.5
10.5.1
10.5.2
10.5.3
10.5.4
10.5.5

Daylight Factor
Components of Daylight Factor
Sky Component TH
Externally Reflected Component TV
Internally Reflected Component TR
Example to Calculating the Daylight Factor

195
195
197
197
198
200

10.6

Influence of Fenestration

205

10.7

Rules for Good Daylighting

207

10.8

Daylight Planning

208

Problems

210

Literature

213

APPENDIX

215

WEATHER DATA

216

PROPERTIES OF BUILDING MATERIALS

223

Symbols
a

Thermal diffusivity

m2/s

Air leakage coefficient

m3/(hmPa2/3)

Area

m2

Thermal effusivity

J/(m2Ks1/2)

Specific heat

J/(kgK)

Specific heat per area

J/(m2K)

Thickness

Energy demand per floor area

MJ/(m2a) or kWh/(m2a)

Total solar energy transmittance

Acceleration of gravity, g = 9.81 m/s2

m/s2

Heat transfer coefficient (surface)

W/(m2K)

Global heat loss coefficient

W/(m2K)

Solar irradiance

W/m2

Air change rate

1/h

Pressure

Pa

Pressure difference

Pa

Power

Heat per area

J/m2

Heat flux

W/m2

Heat (in standards also used as heat per area)

J (or J/m2)

Heat flow

Thermal resistance

m2K/W

Time

Temperature

Time period

Thermal transmittance

W/(m2K)

Velocity

m/s

Volume

m3

Air flow rate

m3/h

x,y,z

Coordinates (in space)

Absorptance

Phase shift

Emissivity

Efficiency or utilization factor

Temperature

Temperature difference

C or K

Thermal conductivity

W/(mK)

Wave length

Heat transfer coefficient (overall)

W/(m2K)

Amplitude damping

Reflectance

Density

kg/m3

Stefan-Boltzmann constant, s = 5.6710-8 W/(m2K4)

Penetration depth

Transmittance

Time constant

s or h

Point thermal transmittance

W/K

Linear thermal transmittance

W/(mK)

Subscripts
a

Air

Convection

External or exterior

Frame

Glazing

Internal or interior

Radiation

Surface

se

Surface external

si

Surface internal

Visible

Window

Symbols Chapter 10 (Daylight)


A

Area

m2

Illuminance

lx

Luminous intensity

cd

Luminance

cd/m2

Luminous efficacy

lm/W

TLQ

Daylight factor

Spectral luminous efficiency (human eye)

Wavelength

m or nm

Transmittance

Visible light transmittance

Luminous flux

lm

Spectral luminous flux

lm/nm

Spectral radiative flux

W/nm

Solid angle

sr

Introduction
Today a great deal of energy is expended to ensure comfortable conditions in the interior space of
buildings. In Switzerland 47% of the total energy demand is generated to operate buildings: heating (35.1%), hot water (5.5%), lighting (3.4%), ventilation and increasingly also cooling (2.7%)
(Source: Swiss Federal Office of Energy, 2008). Also in the European Union the building sector
with 40 % of the energy needs represents the largest energy consumer.
Linked with energy use, especially from fossil fuels, are problems such as pollution, climate change
and the shortage of resources that against the backdrop of the rising world population are becoming increasingly more important.
With sensible design of buildings, especially of the building envelope, the energy demand can today be reduced with reasonable effort by a factor of 5 to 10 compared with older construction. Additionally, in such well-designed structures also the comfort, especially the thermal comfort, is considerably better.
An understanding for the interplay with the outdoor climate and the different energy fluxes in the
building as well as the conditions for thermal comfort is imperative to being able to design, construct and plan energy efficient buildings.

Chapter 1
Heat transfer

Heat Transfer

14

Heat Transfer

With many problems in the natural sciences and technology it is essential to understand and to
quantify heat transport. Heat flow arises because of temperature differences in which the heat
flows from a location of higher temperature to a location of lower temperature. In buildings, especially in the building envelope, large temperature gradients are often present that induce heat flow
(Chapters 4 to 8). The understanding of the mechanisms of heat transfer is also important regarding the thermal comfort of people in interior spaces (Chapter 3) and the energy balance of buildings (Chapter 9). The most important laws of heat transport will therefore be introduced in this
chapter. Heat is transported by different means. A distinction is made between three fundamental
modes of heat transport: conduction, convection (fluid flow) and radiation. Table 1.1 shows in
which mediums these transport modes can occur. As regards radiation the thermal radiation exchange between solid surfaces is of particular importance in building physics.

Heat Transport Mode

Medium

Conduction

solid, liquid, gas

Convection

liquid, gas

Radiation

solid, liquid, gas, vacuum


( system: solid-gas)

Table 1: Heat transport mechanisms


By conduction heat is transported in a material through the mechanical propagation from atomic
and molecular vibrations (collisions). With this no mass transport arises. For example, an iron rod
heated on one side conducts the heat to the colder side. By convection heat is transported through
motion processes (fluid flow); i.e. mass transport takes place. Warm wind, for example, transports
heat in the atmosphere. By thermal radiation thermal energy is transported through electromagnetic radiation. This transport process does not require a medium; i.e. it can also take place in a vacuum. Examples are solar radiation or heat radiation of hot cooktops.

1.1

Conduction

1.1.1 Heat Equation (Fouriers Differential Equation)


In a homogeneous medium a cuboid is considered with side lengths of x, y and z (Fig. 1.1). In
the medium there is a temperature difference. For the sake of clarity only the heat flow in the xdirection is drawn in Figure 1.1. According to Fourier's law of thermal conduction (Jean Baptiste
Joseph Fourier, 1768-1830), the heat flux in the x-direction is:
T
q x =
x

W
2
m

thermal conductivity

W/(mK)

temperature

The thermal conductivity is a material property. The value of the heat flux is higher the larger the
thermal conductivity of the material is and the larger the local change in temperature is. The negative sign indicates that the heat flows from a location of higher temperature to a location of lower
temperature. The heat flux in the y- and z-direction can be determined analogously.

Heat Transfer

15

Fig. 1.1: Geometry to the derivation of the heat equation


The first law of thermodynamics, the conservation of energy, is now applied to the control volume
in Figure 1.1. A balance is established consisting of the heat flow through the surfaces and the
change of stored heat in the volume element.

T T

+
x x y y

T
T
+

= c
t
z z

density

kg/m3

specific heat capacity

J/(kgK)

time

x,y,z

coordinates

With a constant thermal conductivity arises

T
2T 2T 2T

=
+
+
t c x 2 y 2 z 2

This equation is referred to as heat equation or Fouriers differential equation and describes the
spatial and temporal temperature distribution. The value /(c) is denoted thermal diffusivity and
is a measure of the speed of temperature equalization in a medium.
a=

m2

Additionally, the initial conditions belong to the heat equation, given by the temperature distribution
at time t = 0, as well as the boundary conditions. The boundary conditions can be in different
forms. So, for example, the temperature or a certain heat flux can be given at the boundary of the
solution domain.
Table 1.1 gives numerical values for the thermal conductivity, density, specific heat capacity and
the thermal diffusivity of three different materials. The thermal conductivity of good conducting
metals is up to four magnitudes above those of heat insulation materials. In the following three
special cases of the heat equation will be discussed.

Heat Transfer

16

Material

(W/mK)

copper

380

(kg/m3)

c (J/kgK)

a (m /s)
-8

8900

380

1123610

-8

wood (spruce)

0.14

500

2200

1310

mineral wool

0.04

80

600

8310

-8

Tab. 1.1: Thermal conductivity , density , specific heat c and thermal diffusivity a
of different materials.
1.1.2 One-Dimensional, Steady-State Heat Conduction
The simplest heat conduction problem is the case of one-dimensional, steady-state (= timeindependent) heat conduction in a homogeneous layer (Fig. 1.2). The heat equation reduces to

2T
=0
x 2

T
= C1
x

T ( x ) = C2 + C1 x

This implies that a linear temperature profile occurs in the layer.

Fig. 1.2: Temperature profile in a homogeneous wall


With the boundary conditions
T(x = 0) = T0 = C2
as well as Fourier's law of thermal conduction
q
T
= = C1
x

the temperature profile is given by:


T ( x ) = T0

The heat flux through the layer amounts to:

1
q = (T0 T1 ) = (T0 T1 )
d
R
T1 denotes the temperature of the backside of the layer. The value R = d/ is referred to as the
thermal resistance.
The temperature profile in a multi-layered wall is shown in Figure 1.3.

Heat Transfer

17

Fig. 1.3: Temperature profile in a multi-layered wall (itot = 2)


The multi-layered wall can be considered as a series of thermal resistances. For a wall with i layers
the total thermal resistance Rtot is given as follows:
Rtot = Ri =
i

di

The heat flux q through the multi-layered wall amounts to:

1
q =
T0 Ti tot
Rtot

T0 Ti tot denotes the difference of the surface temperatures between all layers. The temperature

change in the i-th layer is

Ti =

Ri
T0 Ti tot
Rtot

The temperature number i is


Ti = T0
i

Ri
T0 Ti tot
Rtot

The temperature Ti can also be easily determined graphically, since in a temperature vs. thermal
resistance diagram the temperature profile can be described by a straight line.
1.1.3 Multi-Dimensional, Steady-State Heat Conduction
In Section 1.1.2 the one-dimensional steady-state heat flow through a wall was examined. In these
cases the temperatures depend only on one coordinate, the coordinate x perpendicular to the layer. This applies to a plane, infinitely long and high wall only. When this assumption is not applicable then more dimensions need to be taken into consideration.
Assuming steady-state conditions the so-called Laplace equation arises from the heat equation.

2T 2T 2T
=0
+
+
x 2 y 2 z 2
In the three-dimensional case the temperature T depends on three coordinates x, y and z:
T = T(x,y,z)

The heat flux q can be regarded as a vector with the three components q x , q y and q z .

Heat Transfer

18

T
T
T
q = grad T =
ex
ey
e
x
y
z z

T
q x =
x

T
q y =
y

T
q z =
z

The components of the heat flux vectors are illustrated in Figure 1.4.

Fig. 1.4: Components of the heat flux vectors in the two-dimensional (left) and in the threedimensional case (right)
For plane problems is T = T(x,y). These temperature field can be represented by lines with constant temperature (= isotherms). The heat flux vector is always perpendicular to an isotherm. The
orthogonal trajectory to the isotherm indicates therefore the path of heat flow (= heat flow line or
adiabat)(Fig. 1.5). An illustration of the two-dimensional heat conduction situation is illustrated in
Fig. 1.6. By analogy with a topographic map the isotherms can be interpreted as contour lines
(constant elevation) and the heat flow lines as the lines, which follow the highest gradient. Twodimensional heat conduction situations will be handled with examples in Section 4.2.

Fig.1.5: Isotherms (T = constant) and heat flow lines.


For some special boundary conditions analytical solutions of the Laplace equation exist. But in
building physics numerical methods are much more important to the solution of multi-dimensional
heat conduction problems.

Heat Transfer

19

Fig. 1.6: Illustration of isotherms and heat flow lines


1.1.4 One-Dimensional, Non-Steady-State Heat Conduction
For one-dimensional, non-steady-state heat conduction the heat equation can be written as follows:

T
2T
=a 2
t
x
Only a few analytical solutions exist for this case. In building physics, therefore, also here the numerical methods are more important. In Chapter 5 this equation will be discussed by means of an
illustrative model.

1.2

Convection

1.2.1 Forced and Free Convection


Convection refers to the transfer of heat by the movement of molecules within liquids and gases.
With this type of heat transfer, mass is transported. In building physics the heat transfer medium is
usually air. One can distinguish between free and forced convection. Free convection, or natural
convection, occurs in liquids or gases if the temperature distribution on the boundary surfaces produces no stable layers. The fluid movement is initiated by density variations due to temperature differences. In an indoor environment free convection can e.g. occur with warm radiators or cold window surfaces. In the case of forced convection the cause of fluid flow is an externally imposed
pressure difference, e.g. from the wind or a fan. The fluid flow can in both cases be laminar (orderly flow with parallel streamlines) or turbulent (chaotic flow). The type of flow laminar or turbulent
influences the heat transfer.
The processes of convective heat transfer are complicated. The most important parameters to the
convective heat flow are:
- physical properties of the fluid (e.g. air)
- temperature difference T btw. wall surface and fluid

Heat Transfer

20

- fluid velocity v
- type of fluid flow (laminar or turbulent Reynolds number Re)
- surface roughness
- heat flow direction (horizontal, vertical up or down)
- geometry
Empirical formulas based on measurements are often used. In this section some useful formulas
will be introduced.
1.2.2 Convection on Surfaces
Firstly, we will consider the heat transfer on inner and exterior surfaces (walls, ceilings, floors). The
heat flux that is caused by convection on a surface amounts to:
W
2
m

q = hc (Ta Ts ) = hc T

hc denotes the heat transfer coefficient (surface) for convection. Ta is the air temperature at a given
distance from the surface and Ts is the surface temperature.
For the case of free convection the heat transfer coefficient with horizontal heat flow (interior wall)
can be approximately described with
hc = 1.31 3 T

W
2
m K

and with vertical heat flow (floors and ceilings of room interiors) with
hc = 1.52 3 T

W
2
m K

Figure 1.7 illustrates the two equations.


hc = 1.52 3 T

Heat transfer coefficient


2

hc (W/(m K))

hc = 1.31 3 T

Temperature difference T

Figure 1.7: Convective heat transfer coefficient as a function of the temperature difference at an
interior surface [1.1].
In the case of forced convection, like with a wind-blown exterior wall, the heat transfer coefficient
can be specified by

Heat Transfer

T
hc = 7.126 v 0.8

273

0 .7

21

W
2
m K

Figure 1.8 shows this graphically and additionally illustrates the influence of the surface roughness.

T
hc = 7.126 v 0.8

273

0 .7

Heat transfer coefficient


2

hc (W/(m K))

Wind speed (m/s or km/h)

Fig. 1.8: Convective heat transfer coefficient on exterior surfaces [1.1]


1.2.3 Convection in Cavities
Figure 1.9 shows a narrow, vertical cavity with isothermal sidewalls at different temperatures. The
fluid heats up on the warm surface, expands and rises up due to the lower density. On the cold
surface it cools down and sinks so that a closed movement results.

Fig. 1.9: Convection in a narrow, vertical cavity (H >> L) with isothermal sidewalls at different
temperatures (Th > Tk).
For example, it is of interest in glazings or in solar collector cavities, whether fluid flow is initiated.
In these cases the convection increases the thermal losses and, therefore, is not desirable.
We consider the case of narrow, vertical cavity with H >> L. Three dimensionless characteristic
values; the Nusselt, Rayleigh and Prandtl numbers; are fundamental to the characterization of the
problem:

Heat Transfer

22

Nu =

hL

()

Ra =

g T L3
a

heat transfer coefficient

cavity width

thermal conductivity of fluid

acceleration of gravity, g = 9.81 m/s2

volumetric thermal expansion coefficient

temperature difference between the walls

kinematic viscosity

thermal diffusivity

()

Pr =

()

Fig. 1.10: Convection in narrow cavities: Nusselt number as a function of Rayleigh number [1.2].
The heat flux through the cavity, caused by convection and conduction, can be expressed with the
Nusselt number as follows:

T
q = Nu
L
For Nu = 1 the heat transport is only by conduction. Figure 1.10 shows the Nusselt number as a
function of the Rayleigh number for Pr = 0.72 (air). The graph shows that the convection is initiated
at a critical Rayleigh number Racrit. The inclination of the cavity to the horizontal plane is plotted as
a curve parameter.
According to [1.3] for narrow horizontal (C = 0.16, n = 0.28) and vertical (C = 0.035, n = 0.38) cavities the following can be applied:
Nu = C Ra n

Heat Transfer

1.3

23

Thermal Radiation

1.3.1 Laws and Characteristics


If heat is transmitted in the form of electromagnetic radiation one speaks of thermal radiation. Every body with a temperature above absolute zero (T > 0 K) emits thermal radiation. The electromagnetic radiation is classified, for example, as ultraviolet, visible or infrared according to the wavelength (Fig. 1.11). Radiation with a wavelength between 0.38 und 0.78 m is visible to the human
eye.

Fig. 1.11: Electromagnetic spectrum


The correlation between the propagation velocity c (the speed of light), wavelength and frequency is:
c=

c0
=

c0 denotes the speed of light in a vacuum (c0 = 2.998108 m/s). The speed of light c is proportionately smaller for mediums with a refractive index n > 1 (glass n 1.5). The frequency is independent of the medium.
Thermal radiation can also be regarded as the flow of photons. A single photon contains the energy
E = h
h denotes Plancks constant (h = 6.62610-34 Js).
When radiation strikes a surface it is either reflected, absorbed or transmitted. The conservation of
energy requires that the sum of the reflectance , absorptance and transmittance is equal to 1:

++=1
In the case of an opaque (= nontransparent) body ( = 0) the equation reduces to

+=1
For an ideal black body s = 1 and s = 0. That signifies that all of the incident radiation is absorbed. The values , and are particularly dependent on the angle of incidence and the wavelength of the radiation. To exemplify a wavelength dependent reflectance, Fig. 1.12 shows () for
different plaster layers. It shows that surfaces that appear white to the human eye must not necessarily be highly reflective also for radiation with longer wavelengths.

Heat Transfer

24

Fig. 1.12: Spectral reflectance of different colored plaster-surfaces [1.4]


The surface roughness determines the spatial distribution of the reflected radiation. If the extent of
roughness is large in comparison to the wavelength of the radiation, the surface reflects diffuse radiation. With reflection for very smooth surfaces the angle of reflection is equal to the incidence
angle.

Fig. 1.13: Direct (left) and diffuse reflection (right) on surface


A black body with a temperature T emits energy according to Plancks law (Max Planck, 1858
1947):
dq s
=
d

C1
C

5 exp( 2 ) 1
T

C1 = 3.7410-16 Wm2
C2 = 0.01439 mK
The wavelength, with which the emitted power spectral density is maximal, can be calculated with
Wiens displacement law (Wilhelm Wien, 1864-1928):

maxT = 2.89810-3 mK
With Wiens displacement law and using the surface temperature of the sun (T 6000 K) one obtains max = 0.5 m, and using a room temperature (T 300 K) a max = 10 m. Figure 1.14 shows
the spectral distribution of radiation from black bodies. In addition the points of maximum emission
are plotted. Figure 1.15 shows the normalized spectral distribution of radiation from black bodies.

Heat Transfer

25

Fig. 1.14: Spectral distribution of radiation from black bodies at different temperatures

Fig. 1.15: Normalized spectral distribution of radiation from black bodies


The total emitted power density from a black surface can be obtained by integration of Plancks
equation:
q s =

dq s
d = s T 4
d

=0

This equation is named Stefan-Boltzmann law (Josef Stefan, 1835-1893 and Ludwig Boltzmann,
1844-1906). s denotes the Stefan-Boltzmann constant (s = 5.6710-8 W/(m2K4)). This law states
that the power emitted is directly proportional to the fourth power of the surface temperature of a
body.

Heat Transfer

26

For gray bodies () = constant < 1 applies and it can be written:


q = q s = s T 4

Kirchhoffs law of thermal radiation (Gustav Robert Kirchhoff, 1824-1887) states, that with a certain
wavelength and temperature the emissivity is equal to the absorptance of a body:

(,T) = (,T)

Wavelength (m)
Fig. 1.16: Spectral emissivities of materials [1.5]

solar
smooth concrete
lime-sand brick
synthetic plaster, white
mineral plaster, gray
fir wood untreated
brick facing, red
aluminum anodized
aluminum chrome-plated
aluminum with Ni, black
Corten steel, raw material
Corten steel, weathered
float glass 6 mm
solar control glass Calorex-A2 (Schott)
insulating glass Comfort (Glaverbel)
glass with SiO2/Au(10 nm)/SiO2
glass with TiO2/Ag(20 nm)/TiO2
glass with Cu(10 nm)/ SiO2(50 nm)

IR (T = 300 K)

(-)

(-)

(-)

0.55
0.60
0.36
0.65
0.44
0.54
0.33
0.33
0.87
0.86
0.86
0.12
0.41
0.34
0.32
0.40
0.34

0
0
0
0
0
0
0
0
0
0
0
0.80
0.36
0.54
0.35
0.21
0.36

0.96
0.96
0.97
0.97
0.92
0.93
0.92
0.07
0.10
0.51
0.92
0.91
0.11
0.12
0.10
0.03
0.05

Table 1.2: Solar absorptance , solar transmittance and emissivity (T = 300 K) of building
materials in the infrared range (IR) [1.6]

Heat Transfer

27

Absorptance and emissivity can be very different for different wavelengths. An example for this are
selective absorbers of solar collectors: high absorptance in the wavelength range of solar radiation
with < 3 m, low emissivity for thermal radiation with > 3 m. Figure 1.16 shows qualitatively
the wavelength dependent emission properties of materials and Table 1.2 gives some numerical
values for solar absorptance, solar transmittance and emissivity of building materials.
As an application example, Figure 1.17 shows the radiation balance of a building surface. I and IIR
denote the intensities of solar and infrared radiation. In addition, not shown in the figure, energy
fluxes occur by thermal conduction in the wall and due to convective losses of the surface to the
exterior.
I
IIR

sTs4

Ts

Fig. 1.17: Radiation balance of a building surface


1.3.2 Radiation Exchange between Surfaces
As an example of radiation exchange between surfaces, the case of two planes of infinite length
and width will be considered here. The plane surfaces are at temperatures T1 and T2, respectively.
It is assumed that both planes are opaque and have wavelength independent emissivities 1 and 2,
respectively. This situation is similar to those in a glazing cavity or the cavity of a solar collector in
the range of infrared radiation. The emitted heat flux E1 from plane number 1 can be calculated as
follows:
E1 = 1 T14

Fig. 1.18: Radiative heat transfer between two parallel planes


From this radiative flow a part is absorbed by plane number 2, the remaining part is reflected and
strikes plane number 1 and so on. Figure 1.18 shows the back and forth reflected radiation produced due to emission by plane number 1.
The absorbed energy in plane number 2 can be added up as follows:
q1 2 = 2 E1 + 2 1 2 E1 + 2 12 22 E1 + ....+ 2 1n 2n E1

Heat Transfer

28

This geometric series can be summarized using the relationship:

n =0

1 x

This leads to:

2 E1
q1 2 =
1 1 2
Analogous is:

1 E 2
q 2 1 =
1 1 2
The net energy flux from plane number 1 to plane number 2 amounts to:

2 E1
E
E 1 E 2
1 2 = 2 1
q12 = q1 2 q 2 1 =
1 1 2 1 1 2
1 1 2
With the definitions for E1 and E2, respectively, and Kirchhoffs law

=1=1
it leads to:

(T14 T24 )
q12 =
1 1
+ 1
1

For the case of black, parallel planes (1 = 2 = 1) it results in:


q12 = (T14 T24 )

With not too large temperature differences the radiation transport can be approximately linearized:

q12 K (T1 T2 )

with

T + T2
K = 4 1

K denotes the transport coefficient for radiation.


It can also be described as:

q12 = C12 T14 T24

with

C12 =

C12 denotes the so-called radiation exchange factor, which is determined by the geometry and
emission characteristics of the surface.
Not always does the total emitted radiation of a body strike another, as in the case of parallel
planes. With the so-called view factor or solid angle factor F, the heat flow due to radiation between two surfaces in arbitrary spatial positions to one another can be calculated:
Q 12 =

1 2 A1 F12
(T14 T24 )
1 ( 1 1 )( 1 2 ) F12 F21

Heat Transfer

29

For the view factor F12 and F21 it gives:


F12 =

cos 1 cos 2
1
dA1 dA2

A1 A1 A2
s2

F21 =

A1
F12
A2

The view factors contain the spatial geometry and are tabulated for different cases (refer to e.g.
[1.7,1.8]). For parallel planes it gives F12 = F21 = 1.

Fig. 1.19: Radiation exchange between two surfaces

1.4

Design Values for the Total Heat Transfer on Surfaces

The specified design values for thermal resistance and heat transfer coefficients in Table 1.3 are
suitable for many calculations, especially for verifications according to SIA 180 [1.9] regarding heat
and moisture protection. These values include the heat transfer due to convection as well as radiation.

application case

location

energy calculations

interior

thermal resistance R = 1/h


heat transfer coefficient h
R = 1/h
h
2
2
0.13 (m K)/W
7.7 W/(m K)

calculation for moisture protection

exterior
building element in ground
exterior climate
building element in ground
interior, top room-half
interior, lower room-half
windows and doors

0.04 (m K)/W
2
0 (m K)/W
2
0.04 (m K)/W
2
0 (m K)/W
2
0.25 (m K)/W
2
0.35 (m K)/W
2
0.15 (m K)/W

25 W/(m K)
2

25 W/(m K)
2

4 W/(m K)
2
2.9 W/(m K)
2
6.7 W/(m K)

Note: The values for the interior sides consider building edges; with covered locations
without air circulation, e.g. behind furniture, higher thermal resistances can yet occur.

Table 1.3: Design values for thermal resistances and heat transfer coefficients according to
SIA 180 [1.9].

Heat Transfer

30

Literature
[1.1]

Zrcher C., Frank T., Bauphysik: Bau und Energie, vdf, Zrich, 2004

[1.2]

Duffie, J. A., Beckman, W. A., Solar Engineering of Thermal Processes, John Wiley &
Sons, New York, 1991

[1.3]

EN 673, Glass in building: Determination of Thermal Transmittance (U value): Calculation


Method, European Committee for Standardization, Brussels, 1997

[1.4]

Sagelsdorff, R., Frank, T., element 29: Wrmeschutz und Energie im Hochbau, Schweizerische Ziegelindustrie, Zrich, 1990

[1.5]

Hauri, H. H., Zrcher, C., Moderne Bauphysik, Verlag der Fachvereine, Zrich, 1984

[1.6]

Zrcher, C., Strahlungsvorgnge an Gebudeoberflchen (Teil 1), 2. Schweizerisches Status-Seminar Wrmeschutz-Forschung im Hochbau, Zrich, 1982

[1.7]

Siegel, R., Howell, J. R., Lohrengel, J., Wrmebertragung durch Strahlung (3 Bnde),
Springer-Verlag, Berlin, 1988

[1.8]

VDI-Wrmeatlas, Verlag des Vereins Deutscher Ingenieure, Dsseldorf, 1984

[1.9]

SIA Norm 180, Wrme- und Feuchteschutz im Hochbau, Schweizerischer Ingenieur- und
Architekten-Verein, Zrich, 1999

[1.10] Grigull., U., Sandner, H., Wrmeleitung, Springer-Verlag, Berlin, 1990


[1.11] Merker, G.P., Konvektive Wrmebertragung, Springer-Verlag, Berlin, 1987
[1.12] Pitts, D. R., Sissom, E. L., Heat Transfer, Schaum's outline series, McGraw-Hill, New York,
1977

Chapter 2
Exterior Climate

Exterior Climate

32

Exterior Climate

The climatic conditions outside of a building are important in many respects. They influence the required power and energy to heat and cool, but also they basically determine the interior climate,
particularly in an unheated or uncooled building. The air exchange that naturally occurs in an interior room will also be influenced by local wind conditions. The amount of daylight that can be used
in building interiors depends on the solar radiation supplied. All these topics will be handled in the
following sections. The exterior climate additionally causes thermal and hygric loads to the building
construction.
The most important climate parameters in building physics are: solar radiation, infrared radiation
(atmosphere, environment), temperature, humidity (see Building Physics III) and wind. Weather
patterns exhibit basically three different components: yearly and daily periodicity as well as a distinct random component. For building planning one therefore mostly uses weather data based on
an average value over many years (e.g. 10 years or more).
In Switzerland, climate data has been systematically recorded since 1864 by the national weather
service (Meteoschweiz). In the early 1980s an automatic measuring network (ANETZ), consisting
of 72 stations, was put into operation. A large number of weather parameters are measured in 10minute intervals. The data is compiled into datasets, e.g. available as hourly average values or as
monthly average values over many years. Also weather data records from more than 7000 stations
all over the world, representing typical years, can be utilized by planners with the commercial database software Meteonorm [2.1]. Long-term monthly average values from Swiss climate data are
also available in various SIA standards [2.2, 2.3].
In view of the trend in buildings that calls for little operating energy, knowledge of local climatic
conditions becomes increasingly significant for building planning. The exterior climate dictates the
outer boundary conditions of the building envelope; the intended use, and accordingly the requirements regarding thermal comfort, determines the inner boundary conditions (see Chapter 3).

2.1

Solar Radiation

2.1.1 Solar Radiation Source


On a human time scale the sun is an inexhaustible energy source. The nuclear fusion processes in
its interior take place at extremely high temperatures of approximately 2107 K. Energy released by
these processes is transported through radiation and convection to the outer layers and from there,
emitted into outer space. The gaseous surface of the sun can be approximated as a black body
with a surface temperature TS of 5760 K. The radiation emitted has the same intensity in all directions (isotropic). The radiative power of the sun can be calculated as follows (Stefan-Boltzmann
law):

PS = 4 rS2 TS4
rS designates the suns radius and s the Stefan-Boltzmann constant (s = 5.6710-8 Wm-2K-4).
Figure 2.1 gives data on the geometry and temperature of the sun and Earth. The diagram is not
shown to scale. If the Earth would have the size of a pea, the sun would in proportion be a sphere
with a diameter of 1 m. The distance between these two celestial bodies in this model would be
108 m.
With the Earths radius rE and a distance from the Earth to the sun R the irradiation on Earth PE
can be calculated as follows:
PE = rS2

rE2
TS4
2
R

Exterior Climate

33

By means of satellite measurements the average solar irradiance can be determined outside the
Earths atmosphere. This so-called solar constant is subject to a variation in the range of 3% due
to the elliptical orbit of the Earth.
I0 = 1367 W/m2

Solar constant:

Sun

Photosphere (T 5760 K)

rS = 0.69510 m

Convection zone (T 130000 K)


Core (T 210 K)
7

11

R = 1.49610

m 1.7%

rE = 6.3710 m
Earth

Fig. 2.1: Geometric relationships (not shown to scale) between the sun und Earth
The total extraterrestrial irradiation PE on the Earth is 1.71017 W (extraterrestrial = outside the atmosphere). From this 1.51018 kWh solar energy reaches the Earths envelope per year. The global
energy demand lies many magnitudes below this value (Tab. 2.1). However, it must be added that
only about half of this radiation energy strikes the Earths surface; only about 30% of the Earths
surface consists of land and also not all of this remaining surface is available solely for energy utilization.
Table 2.2 shows the breakdown of the extraterrestrial irradiation into three different wavelength intervals. In terms of energy, just under half of this radiation is visible. Given by the suns outer surface temperature (T 5760 K) the maximum irradiance is at a wavelength of about 0.5 m, i.e. in
the visible range (Wiens displacement law).
Extraterrestrial Solar Energy
(kWh/a)

Energy Demand 2004


(kWh/a)

18

1110

13

2.410

World

1.510

Switzerland

4.110

13
11

Tab. 2.1: Extraterrestrial irradiation und energy demand


Ultraviolet Radiation
(UV)

Visible Radiation
(VIS)

Infrared Radiation
(IR)

0.30 0.38 m

0.38 0.78 m

0.78 3 m

7%

47%

46%

Wavelength interval
Percentage of energy

Tab. 2.2: Spectral distribution of extraterrestrial solar radiation


2.1.2 Influence of the Atmosphere
The short-wave radiation that is emitted from the sun, with a wavelength between 0.3 und 3 m,
cannot strike the Earths surface unhindered. In the atmosphere the incident radiation is partially
scattered (i.e., the photons course changes) and partially absorbed (i.e., converted to heat).

34

Exterior Climate

Figure 2.2 shows the energy flows of the Earth-atmosphere system. To maintain a steady state,
the power emitted from the Earth into space has to be equal to the power the Earth receives from
the sun. This radiation exchange determines the prevailing temperature on the Earths surface,
which is of course of vital importance for life on Earth. Due to the temperature on Earth, the maximum intensity of emitted long-wave radiation occurs at a wavelength of about 10 m (Wiens displacement law).

Fig. 2.2: Energy flows of the Earth-atmosphere system [2.4]. The solar radiation received is set at
100%. Weather patterns essentially happen in the troposphere, the approximately 10 km thick bottom layer.
The spectral distribution of short-wave solar radiation that reaches the Earths surface, in comparison to long-wave radiation from the atmosphere, is shown in Figure 2.3. The clouds influence the
spectral distribution of the long-wave radiation.

Fig. 2.3: Spectral distribution of short-wave solar radiation and long-wave atmospheric counterradiation [2.5]
The effect of absorption and scattering in the Earths atmosphere on the spectral distribution of the
solar radiation is shown in Figure 2.4. Ozone (O3), water vapor (H2O) and carbon dioxide (CO2)
are the most important atmospheric gases responsible for absorption. Each gas exhibits a characteristic absorption band. Depending on the ratio between the wavelength and the size of the scat-

Exterior Climate

35

tered particles (gas molecules or aerosols, i.e. small particles like dust or water droplets), different
scattering mechanisms occur. Solar radiation is therefore attenuated due to absorption and scattering by crossing the Earths atmosphere.

Fig. 2.4: Spectral distribution of extraterrestrial und terrestrial solar radiation and indications of
wavelength dependent attenuation mechanisms in the atmosphere [2.6]
Depending on the path length of the solar radiation in the atmosphere (incidence angle) the solar
spectrum will change. In building physics however, if a spectral calculation is needed, a constant
solar spectrum will generally be used. A typical spectrum for middle geographic latitudes (Zurich
47) is given in Table 2.3. The spectrum is discretized into 20 wavelength intervals with the same
energy content and is well suited for spectral numerical calculations.

Tab. 2.3: Typical spectral distribution of terrestrial irradiation (AM 2), divided into wavelength intervals of equal energy [2.7].
2.1.3 Direct Radiation on Tilted Surfaces
In this chapter the parameters, which determine the incidence angle of solar radiation on a tilted
surface on Earth will be presented. All formula are valid for the direct (beam) radiation only i.e.
the radiation that by crossing the Earths atmosphere is not scattered and are based on the relative movement of the Earth and sun i.e. on the spatial geometry [2.8].

Exterior Climate

36

As a reminder, in Figure 2.5 a couple of geometrical relationships are presented. The orbit of the
Earth around the sun lies in a plane perpendicular to the plane of the paper. The Earth rotates
once every 24 hours around its own axis, which is inclined 23.45 from its plane of orbit. This inclination is the cause for the seasons. The geographic latitude of a location on the Earths surface
is the angle between the plane of the equator and a line through this location and the Earths center. The circle of longitude , or meridian, passes though both poles. The circle of longitude
through the Greenwich observatory in London was set with a value = 0.

Fig. 2.5: Inclination of the Earths axis in relation to the plane of orbit and the definition of the longitudes and latitudes
In Figure 2.6 an inclined surface is shown, e.g. an exterior wall of a building, and the direct radiation that falls on it. The formulas to calculate the position of the sun and the angle of incidence as
well as the symbols used are listed below.
Solar declination :

= 23.45 sin 360

284 + n

365

( )

Equation of time z (refer to Fig. 2.7):


z = 0.008 cos(t ) 0.122 sin(t ) 0.052 cos(2 t ) 0.157 sin(2 t ) 0.001 cos(3 t ) 0.005 sin(3 t )
360
whereby
t =n
365

Incidence angle :
cos( ) = sin( ) sin( ) cos( ) sin( ) cos( ) sin( ) cos( W ) + cos( ) cos( ) cos( ) cos( )
+ cos( ) sin( ) sin( ) cos( W ) cos( ) + cos( ) sin( ) sin( W ) sin( )

Elevation angle of the sun S:


sin(S ) = sin( ) sin( ) + cos( ) cos( ) cos( )

Azimuth angle of the sun S:

( )

( )

(h )

Exterior Climate

sin( S ) =

w
s

S
n
tS
t0
z

cos( ) sin( )
cos(S )

37

( )

Inclination angle of the wall; horizontal = 0, vertical = 90

Solar declination, angle of the sun at mid-day over the equator plane
-23.45 23.45
Incidence angle in relation to a perpendicular of the wall surface

Geographic longitude of the reference meridian (Zurich 0 = -15)


Geographic longitude of location (Zrich = -8.57)
Azimuth angle between sun and wall = s - w
Azimuth angle of the wall; South = 0, East positive, West negative
Azimuth angle of the sun
Geographic latitude of location; -90 90 (Zurich = 47.38)
Hour angle; = 15(12 - tS); mornings > 0; noon = 0; afternoon < 0
Elevation angle of the sun
Day of the year; 1 n 365
Solar time; tS = t0+ z + t*( 0 - )/15 with t* = 1 h
Local time
Time difference, see equation of time

h
h
h

Fig. 2.6: Definitions of the angles for the calculation of the incidence angle
Figure 2.9 shows the position of the sun (elevation angle S and azimuth angle S) in the course of
a year for an observer in Switzerland ( = 47 und = -8). The sun reaches its highest position on
June 22, the lowest on December 21. In summer the days are longer than the nights, in winter the
reverse applies. Twice a year, on March 21 and on September 23, day and night are equally long.

Exterior Climate

z (min)

38

Day number n
Fig. 2.7: Equation of time
Zenith
Axis of the Earth

Summer
Equinox (21.3/23.9.)
Winter

Elevation angle S ()

Fig. 2.8: Projection of the suns orbit on a hemisphere for a location at latitude

Azimuth angle S ()

Fig. 2.9: Solar position diagram for = 47 and = -8 (Switzerland)


For the calculation of a cast shadow (Fig. 2.10), e.g. from a roof overhang or a balcony, the shading angle is required:

Exterior Climate

tan( ) =

39

tan(S )
cos( S W )

Fig. 2.10: Angle for the calculation of a cast shadow


2.1.4 Global Irradiance
The total irradiance that strikes a particular surface is referred to as the global irradiance:
Global irradiance = Direct (or beam) irradiance + Diffuse irradiance
The global irradiance on a horizontal (Index h) surface is composed of the direct irradiance (Index
b = beam), i.e. the radiation that comes directly from the sun, as well as the diffuse irradiance from
the sky, i.e. the radiation that is scattered at least once as it passes through the Earths atmosphere. Thus:
Ig,h = Ib,h + Id,h

(W/m2)

For the diffuse irradiance it is often assumed that approximately the same amount of energy is irradiated from all directions of the skys hemisphere (isotropic irradiation).
Solar radiation strikes a surface tilted by an angle (Fig. 2.11). The global irradiance on a tilted
surface Ig () is made up of the contribution of the direct irradiance, the diffuse irradiance from the
sky and the global irradiance that is reflected off the ground in the direction of the surface. The tilted surface with an angle faces only one part of the sky. The portion of the hemisphere that the
surface faces amounts to (1+cos())/2 and is referred to as the view factor or the solid angle factor.
The difference to 1 results in (1-cos())/2 and equates to the view factor that is formed by the
ground. The ground reflection lies mostly in the range from 0.2 to 0.4. With snow cover the
ground reflection amounts to about 0.7. The global irradiance on an inclined surface can be calculated as follows:
Ig ( b ) = Ib ( b ) +

1 + cos( b )
1 cos( b )
I d ,h +
I g ,h
2
2

For vertical surfaces ( = 90) it becomes:

I g ,v = Id ,v +

1
1
I d ,h + I g ,h
2
2

Exterior Climate

40

Direct radiation

Diffuse radiation from the sky

Diffuse radiation
from the ground

Fig. 2.11: Direct and diffuse radiation striking a surface with a tilt angle of .
Depending on conditions in the atmosphere, particularly the extent of cloud cover, the percentage
of diffuse to global irradiance will vary considerably. Figure 2.12 gives typical values for this. On
the Swiss Plateau about half of solar energy is irradiated as diffuse radiation in average.

Fig. 2.12: Influence of weather conditions on the global irradiance and the percentage of diffuse irradiance
In Figure 2.13 the solar irradiance in the course of a day for different surface orientations is given.
Figures 2.14 und 2.15 show the monthly global irradiation in a year on surfaces with different orientations. These values are based on the average values from measurements taken at 16 stations in
Germany. The difference between winter and summer is striking. Collectors for active solar systems are often positioned at an angle that approximately corresponds to the latitude of the location.
In the northern hemisphere the south facade is especially suitable to passive use of solar energy
for heating. In comparison to other facade orientations, the irradiation is high in winter and somewhat low (overheating control) in summer. Clearly, the least solar energy is irradiated on the north
facade.

Exterior Climate

41

Fig. 2.13: Global irradiance on surfaces with different orientations in July (50 northern latitude,
big city atmosphere) [2.9]

Fig. 2.14: Global irradiation on southern oriented surfaces with different tilt angles [2.10]

Fig. 2.15: Influence of the orientation of a vertical surface on the irradiated solar energy [2.10]

42

Exterior Climate

Fig. 2.16: Worldwide distribution of global solar irradiation [2.11]

Exterior Climate

43

For different locations on Earth there are major differences in the annual amount of irradiated solar
energy (Tab. 2.4, Fig. 2.16). This results in different potentials with regard to solar energy use.
However, also in the middle latitudes the irradiated solar energy is sufficient to supply a substantial
share of the energy demand in buildings.

Location

Irradiation (MJ/m a)

London
Paris
Zurich
Rome
Cairo
Sahara

3400
4070
4200
6050
7340
8460

Tab. 2.4: Annual global irradiation on a horizontal surface for different locations on Earth
[2.10]

2.2

Air Temperatures

The air temperature at ground level is the result of the energy flows in the Earth-atmosphere system (Chap. 1.1.2). Figure 2.17 shows typical values for the air temperature in Zurich according to
the time of day and month. The highest temperature is reached in July and the lowest in January.
The diurnal cycle exhibits a minimum before sunrise and a maximum in the late afternoon since
there is a time delay between maximum solar irradiance and maximum air temperature due to the
heat storage capacity of the soil.

Fig. 2.17: Typical air temperature in the course of a day for Zurich in different months [2.5]
A condensed presentation of temperatures at a given location is possible by means of cumulative
distributions. The data are thereby added up from the smallest value. The cumulative distribution
function typically has an S-shaped curve and indicates over how many hours per year the temperature is lower, respectively higher, than a given value (Fig. 2.18).

Exterior Climate

44

Number of hours per year with


exterior temperatures above (h)

Number of hours per year with


exterior temperatures below (h)

Temperature (C)

Fig. 2.18: Cumulative distribution of the air temperature for nighttime (A) and daytime (B) [2.12]
In Figure 2.19 the monthly mean values of the air temperature is given for Zurich. In the same figure the monthly irradiation is also plotted. The minima of both values occur at a shift in time due to
the heat storage capacity of the soil. With about the same solar irradiance in the months of November and January, the air temperature in January is significantly lower than in November.

Fig. 2.19: Annual variation of solar irradiance and air temperature for Zurich (Weather data from
[2.3])

Exterior Climate

2.3

45

Soil Temperatures

Temperature (C)

Fig. 2.20 shows the penetration of the seasonal temperature fluctuations in the soil. The annual
mean value of the surface temperature is 13C in this case, which is the very same temperature
that occurs at greater depths. Due to the storage capacity of the soil the temperature fluctuations
that penetrate the ground are delayed (refer to Chap. 5). Thus, for example, the minimum temperature at a depth of 2 m is reached in the month of March. From March through August heat flows into the soil and from September through February the soils heat is released into the air ( temperature gradient).

Annual mean temperature

Depth (m)

Fig. 2.20: Penetration of temperature fluctuations in the soil

2.4

Wind

Height z (m)

The atmospheric air pressure is, in relation to building physics, of secondary importance. The air
pressure decreases with increasing altitude. In Zurich the average air pressure is approximately
95'000 Pa; fluctuations occur in a range of about 3%. These temporal fluctuations occur very
slowly in comparison to fluctuations of the wind-induced pressure.

Fig. 2.21: Wind profile at different surface roughnesses [2.13]

Exterior Climate

46

With regard to the air exchange in buildings, local wind conditions matter. Wind is characterized by
two parameters, wind direction and wind speed (Figures 2.21 und 2.22). They are subject to local
and temporal variations. The wind speed as a function of the height above the ground depends on
the surface roughness. To calculate a wind profile the following empirical power law is often used:

z
v ( z ) = v G
zG

(m/s)

v denotes the wind velocity, z the height above the ground, vG the wind speed at a reference height
zG and the so-called roughness exponent. For built-up areas = 0.4 approximately applies and
= 0.16 for open areas (Fig. 2.21).

Fig. 2.22: Wind direction and speed at Zurich-Kloten [2.5]

2.5

Climate Fluctuations and Changes

In addition to daily, yearly and random fluctuations, exterior air temperatures show also a certain
variability of annual mean values and, additionally, long-term changes (Fig. 2.23).

Temperature (C)

20

15
Lugano

1901-2003 1983-2003
(K/Decade) (K/Decade)
0.117
0.795
Lugano
Geneva
0.161
0.795
Zurich
0.143
0.776
0.642
Davos
0.165

Geneva

10
Zurich
5
Davos
0
1900

1950

2000
Jahr

2050

Fig. 2.23: Measured (1901-2003) [2.14] and predicted [2.15] annual mean air temperatures at four
locations in Switzerland [2.16]
The gradual warming of the global climate that arises from greenhouse gas emissions, especially
from carbon dioxide, is clearly apparent in temperature measurements collected over a number of

Exterior Climate

47

years. In the 20th century, there was a global temperature increase of 0.6 K and an increase of 1.0
to 1.6K in Switzerland. The temperature increase in the last 30 years was particularly pronounced
and larger than what was predicted by all climate models. A global temperature increase from 1.4
to 5.8 is expected in the 21st century [2.15]. The degree of the predicted increase varies with each
scenario (assumed greenhouse gas emissions) and applied climate model.
Considering the long service life of buildings, in Switzerland typically 80 years, the buildings that
we construct today must therefore also function some day in a warmer climate. With higher exterior
temperatures, thermal protection in summertime calls for greater attention.

2.6

Impacts on the Building Envelope

2.6.1 External Heat Transfer Coefficient


The thermal interaction between the external surface of the building envelope and the outside environment depends on both heat transport mechanisms, infrared radiation and convection. The convective heat transfer coefficient increases relatively strongly with increasing air velocity. The following approximate formula can be used:
T
hc = 7.126 v 0.8

273

0 .7

(W/m2K)

Air velocity

m/s

Temperature

As a function of temperature and emissivity, both the building surface and the environment
(ground, neighboring buildings, trees, atmosphere) emit infrared radiation. Under cloudy skies and
with fog the radiation temperature of the atmosphere corresponds somewhat to the outside air
temperature in the vicinity of the building. The total infrared radiation exchange with the environment can be characterized approximately by a radiative heat transfer coefficient from about 4 to 5
W/m2K.
The following value for the total external heat transfer coefficient applies for heat transmission calculations (Chapters 4 und 9, U-value, energy demand) [2.2]:
he = 25 W/m2K (with Re = 1/he = 0.04 m2K/W)
This value contains both the convective and the radiative heat transfer and is based on a conservative assumption. The external heat transfer coefficient is probably slightly overestimated and
thereby also the heat loss through the building element (note: with modern well insulated building
elements, the influence of he on the U-value is however very small).
For building elements below ground it is assumed [2.2]: Re = 0.
2.6.2 Surface Temperature Reduction Due to Infrared Radiation
During a night with a cloudless sky a building surface is in a radiation exchange with higher and
therefore colder air layers than during a night with an overcast sky. This is because the atmosphere becomes partially permeable to infrared radiation with wavelengths of about 10 m and radiates back significantly less than what the air temperature in the building vicinity corresponds to
(Fig. 2.3). As a result the building surface loses additional heat due to infrared radiation (Fig. 2.24).
With good thermal insulation, the heat flow through the wall or glazing is very limited and on a
cloudless night the external surface temperature can drop a few Kelvin below the outside temperature. If the dew point temperature is reached, condensation occurs on the external surface. Particularly with well-insulated glazings, this external surface condensation can be observed sometimes

Exterior Climate

48

on a morning after a cold cloudless night. But also with well-insulated facades and roofs this effect
arises ( algae, fungi). If the surface temperature falls below 0C, frost can even form.

si

interior

Overcast night
sky

exterior

interior

se

si

Cloudless night
sky

exterior

se

qIR
e

Fig. 2.24: Lowering of the external surface temperature se on a cloudless night due to increased
infrared radiative losses q IR to the sky.
2.6.3 Service Life of Exterior Building Elements
The solar irradiation on a building envelope, also on the Swiss Plateau, can cause quite considerable variations in the surface temperatures:
summer: ca. +15C (night) to +80C (day, high solar irradiance)
winter: ca. -10C (night) to +20C (day)
The color of the building external surface also plays a strong role in determining the occurring temperature, due to the different absorptances for solar radiation (Fig. 2.25).

Fig. 2.25: Influence of the facade color on the occurring surface temperature with solar irradiation
These daily temperature fluctuations in combination with exposure to water, air contaminants
(corrosion), wind and ultraviolet radiation greatly stress the exterior elements of the building envelope and thus limit its service life (Tab. 2.5)

Exterior Climate

49

Service life (year)


paint
plaster
window
metal facade
brick facing
concrete

5 - 10
30 - 50
20 - 25
ca. 20
20 - 100
30 - 100

Tab. 2.5: Service life of exterior building elements [2.9]

Problems
Problem 1: Shading on a Window
A window is set into a Southwest orientated facade with a recess of 20 cm. What percentage of the
window with the dimensions 1.4 m x 1.4 m will be in shade from the wall protrusion in summer
(June 23) at 15.30 h (standard time)? What other objects can cast a shadow on a building faade?
Name some examples.

Problem 2: Wind Velocity und Solar Irradiance on a High-Rise Building


In Zurich a high-rise building is planned for with a square footprint, a side length s = 30 m and a
height h = 120 m with 40 stories each having 900 m2 floor area.
a.) What wind speed can be expected at a height z2 at the top floor, if at a height z1 = 10 m a velocity of v1 = 50 km/h is measured? Use the power law below to describe the wind velocity profile for your assessment und take a large surface roughness ( = 0.4).

z
v 2 = v1 2
z1

b.) How much solar energy will be irradiated on the building envelope faades N/E/S/W and horizontal roof per year (MJ/a)? Use the weather data for the city of Zurich (see Appendix).
c.) Apply the results from b.) to obtain the irradiated solar energy per square meter floor area.
d.) Assume that a similar building is planned for in the vicinity, in which however the side lengths,
height and number of floors n are modified by a factor . How do the results for this building
differ from those in problem c.)?

Literature
[2.1]
[2.2]
[2.3]
[2.4]

Meteonorm 6.0 (Edition 2007), Datenbanksoftware, Meteotest, Bern;


http://www.meteonorm.com
SIA 180, Wrme- und Feuchteschutz im Hochbau, Schweizerischer Ingenieur- und Architekten-Verein, Zrich, 1999
SIA 381/2, Klimadaten zu Empfehlung 380/1 Energie im Hochbau, Schweizerischer Ingenieur- und Architekten-Verein, Zrich, 1991
Volz A., Studie ber die Auswirkungen von Kohlendioxid-Emissionen auf das Klima, Bericht
der KFA Jlich Nr. 1877, 1983

50

Exterior Climate

[2.5]

Zrcher C., Frank T., Bauphysik: Bau und Energie, vdf, Zrich, 2004

[2.6]

Iqbal M., An Introduction to Solar Radiation, Academic Press, Toronto, 1983

[2.7]

Wiebelt J.A., Henderson J.B., Selected Ordinates for Total Solar Radiation Property Evaluation from Spectral Data, Trans. of the ASME, J. of Heat Transfer, 101(101), 1979

[2.8]

Duffie J.A., Beckman W.A., Solar Engineering of Thermal Processes, John Wiley & Sons,
New York, 1991

[2.9]

Keller B., Bauphysik: Die Energetik des Gebudes, Vorlesungsskript ETH, Zrich, 2006

[2.10] Goetzberger A., Wittwer V., Sonnenenergie: Thermische Nutzung, Teubner, Stuttgart, 1989
[2.11] Energieatlas GmbH, CH-4142 Mnchenstein, 2005; Datenquelle: Meteonorm 4.0, Meteotest, Bern
[2.12] Baumgartner T., Steinemann U., Geiger W., Meteodaten fr die Haustechnik, SIA Dokumentation D012, 1987
[2.13] Moor H., Physikalische Grundlagen der Gebudeaerodynamik im Hinblick auf die Berechnung des Luftaustausches, EMPA Dbendorf, 1987
[2.14] Begert M., Seiz G., Schlegel T., Musa M., Baudraz G., Moesch M., Homogenization of
measured time series of climatic parameters in Switzerland and computation of norm values 1961-1990. Final project report NORM90, MeteoSwiss, Zurich, 2003
[2.15] Intergovernmental Panel on Climate Change: http://www.ipcc.ch. Retrieved October 2004.
[2.16] Christenson M., Manz H., Gyalistras D., Climate warming impact on degree-days and building energy demand in Switzerland, Energy Conversion and Management, Vol. 47, 2006,
671-686

Chapter 3
Thermal Comfort

high temperature

low temperature

Figure: Infrared Image of a Person

Thermal Comfort

52

Thermal Comfort

Whether a person feels comfortable in an indoor environment depends not only on the thermal
conditions of the room, but also on the air quality, the acoustics (noise level, reverberation, etc.),
visual (availability of natural light, glare, etc.) and additional factors (occupancy rate, furnishings,
aesthetics, etc.). The thermal comfort is however an important aspect of total comfort.
People in indoor environments are connected thermally with the building structure. The heat
transport processes between a human and the room, as well as the conditions with which people
feel satisfied with their sensation to heat, shall thus be presented in the following. Thermal comfort
is on the one hand determined by the building envelope, and on the other hand by the HVAC facilities. The mechanical equipment can be regarded as supplementary measures to balance out the
inadequacy of the building envelope. Because the operation of the mechanical equipment requires
energy, the goal must be to design the building envelope to guarantee good comfort with as little
mechanical equipment as possible. From this point of view the resulting requirements on the building envelope shall be introduced.

3.1

Human and Interior Space

Since a person is a source of heat in an indoor environment, the air in the immediate vicinity of the
human body will be heated. The air consequently expands and a buoyancy flow arises that envelops the person. About 60 to 100 m3/h of air is transported above from below. The boundary layer
of the rising air (Fig. 3.1) has a thickness of approximately 10 to 15 cm. Comparable updrafts also
arise from other heat sources in the room, as for example from a radiator, a computer or a lamp.

Temperature
Heat flow
Velocity v

Boundary layer

Distance x

Fig. 3.1: Temperature- und air velocity profile at a warm surface

Buoyancy
flow
Downdraft

Radiation
exchange

Fig. 3.2: Convective und radiative heat flows between a human and a room

Thermal Comfort

53

At cold surfaces, especially windows in winter, downdrafts (cold sinking air) develop. Humans in an
interior find themselves in a landscape of boundary layer currents that they themselves influence.
These landscapes can also be influenced naturally by the momentum of the entering air current;
from air inlets of mechanical ventilation systems or from open windows. Additionally, radiative heat
exchange takes place between a person and the environment (Fig. 3.2). With an interior temperature of T 300 K, this radiation exchange occurs in the infrared region with a maximum intensity at
a wavelength of about 10 m (Wiens displacement law).

3.2

Heat Balance of Humans

Humans belong to the organisms that maintain a constant internal body temperature also with
changing temperature in the environment. This results from the bodys own thermoregulatory system. The so-called core body temperature is about 37C. The body exterior is subject to large temperature variations. Figure 3.3 shows the temperature distribution in the human body.

Fig. 3.3: Temperature distribution in the human body with different ambient conditions [3.1]
The food that humans ingest contains energy that is converted into heat by a biochemical process.
This exothermic reaction in the human body is often referred to as metabolism. The rate of this
heat production in the human body is notably dependent on the level of physical activity. Table 3.1
shows the metabolic rate in relation to the activity. An adult can, e.g. during a foot race, reach a
temporary metabolic rate of more than 800 Watt. The unit met serves as a description of the degree of activity (1 met = 58 W/m2 = 104 W/Person = seated, relaxed).
The average heat flux through the body surface of a relaxed seated person is given as:
W
Q 104W
q = =
= 58 2
2
A 1 .8 m
m

The heat flux varies depending on the body part [3.2]:


head: ca. 115 W/m2
hand: ca. 75 W/m2
foot-sole: up to 145 W/m2

Thermal Comfort

54

As already mentioned above, the temperature in the internal body must be kept at a constant temperature. It thus follows that the sum of the heat output and enthalpy flow plus the mechanical output must be equal to the metabolic output (conservation of energy):
Metabolic output = Heat- and enthalpy flow + Mechanical output
This energy flow balance somewhat dramatized could also be called the equation of life and
death. The mechanical output is generally small in comparison with the other values and can be
neglected for the usual indoor activities. The heat flow is a combination of the heat loss of the body
through convection to the ambient air, plus a contribution produced through the infrared radiation
exchange between a person and the enclosing surfaces (walls including windows, floors and ceilings). There where the body is in direct contact with the surroundings e.g. on the floor or if the legs,
seat or back are in contact with a chair, the heat can additionally flow by conduction. Enthalpy flow
occurs because the evaporation of sweat extracts body heat. The exhalation of moist breath also
adds to enthalpy flow (Figures 3.4 und 3.5).
With high ambient temperatures (e.g. in a sauna) or with strong physical labor, not enough energy
can be lead away by radiation and convection. Under these conditions heat must be given off primarily by evaporation of water (Fig. 3.6); assuming that the air moisture is not too high.
Activity

Metabolic rates

Reclining, asleep
Seated, relaxed
Sedentary activity (office, dwelling, school, laboratory)
Standing, relaxed
Light activity, standing (shopping, light industry, laboratory)
Medium activity, standing (domestic work, machine work)
Walking (4 km/h)
Heavy activity (heavy industry)
Walking (5 km/h)
Running (10 km/h)

met

W/m *

W/Person**

0.8
1.0
1.2
1.2
1.6
2.0
2.8
3.0
3.4
8.0

46
58
70
70
93
116
162
174
197
464

83
104
126
126
167
209
292
313
354
834

* based on the body surface area


** valid for one person of 1.8 m2 body-surface area (e.g. height 1.7 m, weight 69 kg)

Table 3.1: Heat production with different activities [3.3]

Fig. 3.4: Heat loss from the skin due to radiation, convection, and evaporation [3.1]

Thermal Comfort

55

Fig. 3.5: The heat loss mechanisms of a partially clothed body [3.4]

Fig. 3.6: Heat loss with different ambient temperatures [3.4]


2

Type of Clothing

clo

m K/W

Naked, standing
Underpants, bathing suit
Typical tropical clothing: underpants, shirt/blouse with short sleeves and open collar, shorts, light socks and sandals
Light summer clothing: underpants, shirt/blouse with short sleeves and open collar, light-weight trousers or skirt, light socks and shoes
Light work clothing: underwear, shirt/blouse with short sleeves and open collar,
work trousers, socks and shoes
House clothing in winter: underwear, shirt/blouse with long sleeves, sweater with
long sleeves, trousers or skirt, thick socks and shoes
Traditional winter clothing: long-underwear, shirt with long sleeves, suit with trousers, jacket and vest or dress, thick socks and shoes
Warm winter clothing

0.0
0.1
0.3

0.0
0.015
0.045

0.5

0.08

0.7

0.11

1.0

0.155

1.5

0.23

3.0

0.45

Table 3.2: Thermal resistances of different clothing ensembles [3.3]

Thermal Comfort

56

Clothing acts as a resistance to both the heat flow and the vapor diffusion flow. The thermal insulation values of different clothing ensembles are shown in Table 3.2. A typical clothing ensemble in
winter for an interior space is designated as 1 clo, the abbreviation for cloth.
Mass
Volume
Surface area
Core body temperature
Skin temperature
Inhaled airflow rate
Breath rate
Pulse
H2O-Production
CO2-Production

70 kg
3
70 dm
2
1.8 m
37C
34C
3
0.5 m /h
-1
16 min
-1
60 - 80 min
40 g/h
15 - 20 l/min

Table 3.3: Data on the human body

3.3

Requirements for Thermal Comfort

The requirements for thermal comfort, and respectively a persons discomfort in a room, can be divided into two areas of influence:
Influence of the room:

air temperature
surface temperatures of the enclosing areas
air movement (velocity, turbulence intensity, direction)
relative air humidity

Influence of people:
physical activity (heat production)
clothing (thermal insulation)
physiological condition
Not everyone is sensitive to the same things in the same way. All predictions are therefore statistical. The room conditions that apply as optimal are those that feel comfortable for most of the occupants. In the SIA standard 180 [3.3] it is stipulated that 80% of the occupants, provided that they
are engaged in normal activity and dressed appropriately for the season, should find the room
conditions comfortable. A PPD index establishes a Predicted Percentage of Dissatisfied [3.5]. The
SIA standard 180 thereby requires a PPD < 20%.
As an approximation, the perceived room temperature, which is designated as the operative room
temperature o, can be calculated as the mean air temperature a plus the mean radiant temperature r , where r is assumed as the mean surface temperature of the enclosures (walls, ceiling
and floor):

o =

a + r
2

where r =

1
ktot

Ak

ktot

Ak sk

(C)

k =1

k =1

Ak

area of the k-th enclosure surface

m2

sk

temperature of the k-th surface area

Thermal Comfort

57

From this relationship it follows that e.g. lower surface temperatures can be compensated by higher air temperatures. The temperature difference should however not be larger than about 1.5 to 3 K
[3.6].

Specific metabolic rate (met)

In Figure 3.7 the optimal room temperature as a function of physical activity and clothing can be
determined. So, for example, with a sedentary activity and normal daily wear clothing in winter the
optimal room temperature o is at 21.5C with a tolerance range of 2C. Figure 3.7 also gives the
temperature tolerance, within which a PPD < 10% applies. According to [3.3] only a PPD < 20% is
required, therefore a range is included in Figure 3.7 to account for comfort losses not covered. In
the following chapter these additional comfort parameters will be discussed in greater detail. In Table 3.4 the comfort requirements specified in [3.3] are displayed.

Thermal resistance of clothing (clo)

Fig. 3.7: Optimal operative temperature o as a function of activity and clothing [2.3]. The shaded
areas indicate the temperature range inside which the condition PPD < 10% is fulfilled.

Parameter

Winter
(clothing 1 clo and
activity 1.2 met)

Summer
(clothing 0.5 clo and
activity 1.2 met)

- room temperature
- air temperature difference
(0.1 m to 1.1 m above floor)
- floor temperature
- max. asymmetry of
radiant temperature

19C 24C
<3K

23.5C 26.5C
<3K

19C 26C
- heated ceiling 4 K
- cold walls 10 K
- warm walls 20 K
< 0.15 m/s

n/a
- cold walls 10 K
- cold ceiling 13 K
- warm ceiling 5 K
< 0.2 m/s

- air velocity

Table 3.4: Comfort requirements according to [3.3]

Thermal Comfort

58

3.4

Detailed Comfort Analysis

Based on numerous experiments on subjects, P. Ole Fanger (1934-2006) [3.5, 3.7, 3.8, 3.9, 3.10]
developed a model that can predict the thermal comfort of people under the application of physical
and physiological quantities, in indoor environments with a moderate climate. Based on a global,
steady-state energy balance of the human body, refer to Chapter 3.2, an average vote PMV (Predicted Mean Vote) can thereby be predicted. The predicted percentage of dissatisfied PPD is then
treated as a function of the average PMV value. That is, the PMV- and PPD-Indexes describe the
hot and cold discomfort of the whole body.
Discomfort can however also result from unwanted cooling or heating of one particular part of the
body. One then speaks about local discomfort. Causes for this can be drafts, radiant temperature
asymmetry, or high or low floor surface temperatures.

3.4.1 Global Comfort Criteria


The predicted mean vote PMV [3.5, 3.7] is dependent on both the influences of the environment
(air temperature and surface temperatures, air movement and humidity) and on the person (physical activity, clothing). The PMV is an index that predicts, with the help of a 7-point thermal sensation scale, the mean value of the votes of a large group of persons (Table 3.5).
PMV

-3

-2

-1

+1

+2

+3

Assessment

cold

cool

slightly
cool

neutral

slightly
warm

warm

hot

Table 3.5: Thermal sensation scale after Fanger [3.5]


The predicted mean vote PMV is given by the equation [3.5]:

){

PMV = 0.303 e 0.036 M + 0.028 M W 3.05 10 3 [5733 6.99 (M W ) p ]

0.42 [M W 58.15 ] 1.7 10 5 M (5867 p ) 0.0014 M (34 a ) 3.96 10 8 fcl

4
4
( cl + 273 ) r + 273 fcl hc ( cl a )}

M
W
Rcl

a
r
p

metabolic rate, per square meter of body surface area


external work (typically W = 0), per square meter of body surface area
thermal resistance of clothing
air temperature
mean radiant temperature
partial water vapor pressure

Surface temperature of clothing:

cl = 35.7 0.028(M W ) R cl {3.96 10 8 fcl ( cl + 273 ) r + 273


4

) ]+ f h (
4

cl

cl

W/m2
W/m2
m2K/W
C
C
Pa

a )}

Convective heat transfer coefficient:

0.25
hc = max 2.38 ( cl a ) , 12.1 v

(W/m2K)

m/s

relative air velocity

Ratio of persons surface area while clothed to while nude:

(C)

Thermal Comfort

1.00 + 1.290 Rcl ,


fcl =

1.05 + 0.645 Rcl ,

fr

Rcl < 0.078m 2 K / W

fr

Rcl > 0.078m 2 K / W

59

Mean radiant temperature:


k tot

r = FPk sk
k =1

FPk

sk

k tot

(C)

where

FPk

=1

k =1

radiant view factor between a person P and the surface k


temperature of the k-th surface

The radiation exchange between a person and the environment is better described with respect to
the spatial geometry by means of the view factor (solid angle) than with the mean value of the surface temperatures of the enclosing surfaces. Figure 3.8 presents diagrams to determine the view
factor FPk between a person and the surrounding surfaces.

Fig. 3.8: View factor between a seated person and a rectangular surface [3.7]

Thermal Comfort

60

The predicted percentage of dissatisfied can be calculated by an empirical formula [3.5]:


PPD = 100 95 e (0.03353 PMV

+ 0.2179 PMV 2

This equation is shown graphically in Figure 3.9. It should be noted that even under optimal environmental conditions 5% of the people will still be dissatisfied.

Fig. 3.9: Predicted percentage of dissatisfied PPD in relation to the predicted mean vote PMV [3.5]

3.4.2 Local Comfort Criteria


Local discomfort can occur from the following:
air movement
asymmetrical radiant fluxes
floor temperature
3.4.2.1 Draft
Draft is an unwanted local cooling of the human body by air movement. In addition to air temperature and velocity, the turbulence intensity also influences the sensation to draft. The greater the air
velocity is as well as its fluctuations (turbulence), the higher is the convective heat transfer coefficient and with that, the cooling effect of the air on the skin.
The predicted percentage of dissatisfied with respect to draft, i.e., the Draft Rating DR can be calculated as follows [3.3, 3.5]:
DR = (34 a ) (v 0.05 )

0.62

(0.37 v Tu + 3.14 )

if v < 0.05 m/s, then v = 0.05 m/s


if DR > 100%, then DR = 100%

(%)

Thermal Comfort

Tu

61

local turbulence intensity, ratio of the standard deviation of


the local air velocity to the local mean air velocity

The model to calculate the draft rating is based on studies comprising 150 subjects exposed to air
temperatures of 20C to 26C, mean air velocities of 0.05 m/s to 0.4 m/s and turbulence intensities
of 0% to 70%. The model applies to people performing light, mainly sedentary activity, with a thermal sensation for the whole body close to neutral (Fig. 3.10).
In different studies it was also shown that the exposure time, the extent of physical activity as well
as the direction of airflow, was significant to the sensation to draft [3.11, 3.12, 3.13, 3.14]. These
factors were however not considered in the model described above.

Fig. 3.10: Combination of air velocity and turbulence intensity with a DR < 20% [3.3]

3.4.2.2 Radiant Temperature Asymmetry


An asymmetrical radiation exchange between a person and the environment e.g. caused by cold or
warm window surfaces or a heated ceiling, and accordingly a cooled ceiling, can cause discomfort,
also if the global energy balance corresponds to a comfortable state.
The asymmetry of the radiant fluxes can be described by the difference of the mean surface temperatures of both half spaces:

r = r 1 r 2

r1
r 2

(C )

Mean surface temperature of half space 1

Mean surface temperature of half space 2

Figure 3.11 shows the influence of the radiant temperature asymmetry r on the satisfaction.

3.4.2.3 Floor Temperature


Too low or too high floor temperatures can also cause local discomfort (Fig. 3.12).

Thermal Comfort

62

Fig. 3.11: Predicted percentage of dissatisfied with radiant temperature asymmetry [3.8]

Fig. 3.12: Influence of the floor temperature on the predicted percentage of dissatisfied with normal
shoes for seated and standing persons [3.4]

3.5

Cold downdraft

A cold downdraft occurs at vertical surfaces especially at window glazings in winter because the
thermal resistance of the building envelope and the resulting internal surface temperature is in
general considerably smaller here than on the neighboring opaque elements. This descending airflow (Fig. 3.13) can cause discomfort. The maximum air velocity as a function of the distance x
from the cold vertical surface can be approximated by the following empirical formula determined
with an air temperature of 20C [3.15, 3.16, 3.17]:
v max = 0.083 H
0.143
v max =
H
x + 1.32

x < 0.4 m

v max = 0.043 H

x>2m

0.4 m x 2 m

distance from the cold surface


temperature difference between the air and cold surface
height of the cold surface

m
K
m

Thermal Comfort

63

The maximum air velocity at the bottom of the cold vertical surface is usually less notable than the
maximum velocity in the habitation area (x 1 m). The temperature difference between the air and
the cold surface amounts to

U
( i e )
hi

Should a specific air velocity not be exceeded for comfort reasons, e.g. vmax < 0.15 m/s, a required
thermal transmittance value U can be determined for a given value of glazing height H, distance x,
internal heat transfer coefficient hi and interior and exterior temperature, i and e. Low thermal
transmittance values are not only advantageous from an energy point of view, but they also have a
positive effect on thermal comfort due to the small temperature difference between the air and internal surfaces of the building envelope.

Fig. 3.13: Cold downdraft from vertical surfaces

Fig. 3.14: Maximum air velocity from cold downdrafts as a function of distance, temperature difference and height

Thermal Comfort

64

3.6

Comfort Measurements

With comfort measuring instruments relevant quantities in buildings can be determined by measurements. Complaints can be evaluated relatively quickly and objectively as to whether an adequate thermal comfort is provided for, or not. Figure 3.15 shows such an instrument with sensors to
measure the air temperature, air velocity and turbulence intensity, air humidity as well as the
asymmetry of the radiant temperature. The sensor to measure the surface temperature is not
shown in this photo. By applying a software as well as the Fanger formulae to the measured data,
the PMV, PPD und DR values can be determined.

air humidity

air velocity and turbulence intensity

air temperature
asymmetry of
radiant temperature

Fig. 3.15: Sensors for comfort assessments mounted on a tripod

3.7

Adaptive Comfort

Humans have the ability to adapt to the changes in their thermal surroundings in order to reduce a
possible discomfort. This can result not only by changing the clothing and the physical activity, but
also from cold or hot food and drinks as well as air movement (fan). Recent studies indicate that
this ability to adapt plays a role in the sensation of comfort.

Fig. 3.16: Acceptable operative room temperature in naturally ventilated rooms [3.19]

Thermal Comfort

65

Models for adaptive comfort are often based on the assumption that weather conditions codetermine the sensation of comfort [3.18]. Thus the higher the operative room temperature that is
still felt as comfortable, the higher is the average outdoor air temperature (Fig. 3.16). For user satisfaction, it is mostly beneficial to give the user the possibility to set the interior room climate themselves, within fixed constraints, e.g. to allow one to open the windows in summer. Surveys have
shown that using this approach a larger range of operative room temperature is still accepted.
Thereby the potential costs of the HVAC system and the energy use in a building is reduced.

3.8

Implications

To assure good thermal comfort, the following goals should be aimed for:
appropriate, average air temperature (winter ca. 19C to 24C, summer ca. 23.5C to 26.5C,
max. 28C) and minor levels of temperature stratifications
balanced, comfortable surface temperature (minor radiant temperature asymmetry)
low air velocities (v < 0.15 m/s) und low turbulence intensities
With mechanical ventilation or in air-conditioned rooms, the positioning and the properties of the air
supplies as well as the supply air conditions (temperature, velocity, turbulence intensity) are particularly significant for the thermal comfort.
The goal must however be that the building is designed so that as little mechanical equipment
measures as possible are required. With respect to thermal comfort it is imperative to aim for:
1. a very well and uniformly insulated building envelope
2. a high airtightness of the building envelope
In addition to the advantages of a simple HVAC system, a well-insulated and airtight building envelope brings about the most important secondary effect; that much less energy is required to keep
the building comfortable (Chapter 9).

Problems
Problem 1: Human Sensation of Heat and the Influence of Clothing
Under normal conditions a relaxed seated person produces a thermal output of about 104 W. With
a body surface area of about 1.8 m2 this amounts to a heat flow density q = 58.0 W/m2 = 1.0 met
(Table 3.1). The core temperature of the body is at 37C, the skin surface temperature is at 34C.
a) Therefore, how large is the thermal resistance between the body core and skin surface?
b) The heat transfer coefficient of the clothing surface to the surroundings amounts to h = 10
W/m2K. How large is the corresponding heat transfer resistance?
c) The ambient temperature is at 15C. How large must the total resistance between the body core
and the surroundings therefore be, so that the upper given value of the heat flux density is not
exceeded?
d) How large must the resistance of the clothing be, in order to reach this value?
e) Which clothing, with respect to the amount of "clo's", does this correspond to?

Problem 2: Optimal Room Temperature in a Laboratory


The workers in a laboratory wear clothes that correspond to house clothing in winter, with respect
to thermal insulation. One can assume that standing light activity is performed. How large is the
optimal room temperature as well as the tolerance range of room temperature, within which the
condition PPD < 10 % is fulfilled?

Thermal Comfort

66

Problem 3: Cold Air Downdraft near Windows


A full-height window in a room (H = 2.2 m) shall be designed so that in winter (e = -5C und i =
20C) no heating device (e.g. radiator) is required to limit the cold air downdraft, i.e. at a distance
of 1 m from the window the air velocity should not be larger than 0.15 m/s. The heat transfer coefficient on the internal side is 7.7 W/m2K. Determine the required thermal transmittance U for the
window.

Problem 4: Optimal Room Temperature in a Thermal Bath and a Fitness Room


In a resting room of a thermal bath the people typically wear a bathing suit with a bathrobe (thermal
insulation about 0.04 m2K/W). The physical activity can be taken as seated, relaxed.
In the adjacent fitness room the room temperature is set at 24C. The people give off based on
their physical activity on the average about 300 W and wear clothes that approximately corresponds to 0.4 clo.
a.) What room temperature is optimal for the resting room of the thermal bath?
b.) What range of room temperature would you recommend for the resting room?
c.) What optimal room temperature would you recommend for the fitness room?
d.) Which climate assessment (too cold/neutral/too warm) do you expect from the majority of users of the fitness room?
e.) Which heat dissipation mechanism does the body of the fitness room user react with in this
situation?

Literature
[3.1]

Silbernagel S., Despopoulos A., Taschenatlas der Physiologie, Georg Thieme Verlag Stuttgart, 1983

[3.2]

Keller B., Bauphysik: Die Energetik des Gebudes, Vorlesungsskript ETH, Zrich, 2006

[3.3]

SIA 180, Wrme- und Feuchteschutz im Hochbau, Schweizerischer Ingenieur- und Architekten-Verein, Zrich, 1999

[3.4]

Zrcher C., Frank T., Bauphysik: Bau und Energie, vdf, Zrich, 2004

[3.5]

EN ISO 7730, Ergonomie der thermischen Umgebung, Europisches Komitee fr Normung,


Brssel, 2005

[3.6]

Willems W.M., Schild K., Dinter S., Vieweg Handbuch Bauphysik: Teil 1, Vieweg & Sohn
Verlag, Wiesbaden, 2006

[3.7]

Fanger P.O., Thermal Comfort, Danish Technical Press, Copenhagen, 1970

[3.8]

Fanger P.O., Radiation and Discomfort, ASHRAE Journal, 28(2), 33-34, 1986

[3.9]

Fanger P.O., Melikov A.K., Hanzawa H. and Ring J., Air Turbulence and Sensation of
Draught, Energy and Buildings, 12, 21-39,1988

[3.10] Fanger P.O., Melikov A.K., Hanzawa H. and Ring J., Turbulence and Draft, ASHRAE Journal, 31(4), 18-25, 1989
[3.11] Griefhahn B., Assessment of draught at workplaces, Fb 828, Dortmund, Schriftenreihe der
Bundesanstalt fr Arbeitsschutz und Arbeitsmedizin, 1999
[3.12] Toftum J., Zhou G., Melikov A., Air flow direction and human sensitivity to draught, In: Proc.
of CLIMA 2000, Brussels, 1997

Thermal Comfort

67

[3.13] Toftum J., Nielsen R., Draught sensitivity is influenced by general thermal sensation, International Journal of Industrial Ergonomics, 1996, 18(4), 295-305
[3.14] Toftum J., Nielsen R., Impact of metabolic rate on human response to air movements during work in cool environments, Int. J. of Industrial Ergonomics, 1996, 18(4), 307-316
[3.15] Heiselberg P., Draught risk from cold vertical surfaces, Building and Environment, 1994,
29(3), 297-301
[3.16] Heiselberg P., Overby H., Bjorn E., Energy-efficient measures to avoid downdraft from large
glazed faades, ASHRAE Transactions, 1995, 1127-1135
[3.17] Manz H., Frank T., Analysis of thermal comfort near cold vertical surfaces by means of
computational fluid dynamics, Indoor and Built Environment, 13 (3), 2004, 233-242
[3.18] ASHRAE Fundamentals Handbook, Chapter 8: Thermal Comfort, American Society of
Heating, Refrigerating, and Air-Conditioning Engineers (ASHRAE), 2001
[3.19] ASHRAE Standard 55, Thermal Environmental Conditions for Human Occupancy, American Society of Heating, Refrigerating, and Air-Conditioning Engineers (ASHRAE), 2004

Chapter 4
Steady-State Thermal Transmission

Steady-State Thermal Transmission

70

Steady-State Thermal Transmission

In this chapter the heat flow through a building wall, floor or roof with constant boundary conditions
with respect to time will be handled. The heat transmission is in this case referred to as steadystate. Non-steady-state thermal transmission will be presented in chapter 5. Both states of transmission are characterized in Table 4.1.

temperature difference

heat flow

thermal condition

steady-state

constant

constant

flow equilibrium

non-steady-state

not constant

not constant

time-varying

Tab. 4.1: Characteristics of steady-state and non-steady-state

4.1

Standard Cross Section

The heat flow through a wall, floor or roof structure is determined by the thermal resistances of the
individual material layers, the heat transfer resistance between the wall surfaces and the interior
and exterior air, as well as the temperature difference between the interior and exterior.
In a steady-state condition thermal conductivity is the material quantity that dictates the heat flow in
the material. Figure 4.1 gives the thermal conductivity of different materials plotted against their
density (cp. Appendix).

Fig. 4.1: Thermal conductivity plotted against density [4.1]


The heat transfer coefficient h between the wall surface and the air is comprised of contributions
from both radiation and convection (cp. chapter 1):
h = hr + hc

(W/(m2K))

Steady-State Thermal Transmission

71

Values for the heat transfer coefficient h are given in chapter 1.4. Figure 4.2 shows the resistance
to the heat flow when crossing through an exterior wall. In this figure an analogy to electrical engineering is shown that illustrates a thermal equivalent circuit diagram with resistances.

se

si

dn

1
he

1
hi

1
U

R1

Re

Rntot
R

Re

Ri
Ri

Rtot
Fig. 4.2: Thermal resistances of a multi-layered wall
The exterior heat transfer resistance Re, the interior heat transfer resistance Ri and the thermal resistance R through the layers of construction material can be calculated as follows:
Re =

1
, R=
he

ntot

dn

n =1

, Ri =

m2 K

1
hi

dn denotes the thickness of the n-th layer. The total thermal resistance from the inside to the outside adds up to:
R tot = R e + R + R i =

1 ntot d n 1
+
+
he n =1 n hi

m2 K

The thermal transmittance U is the reciprocal value of the total thermal resistance:
U=

1
1
=
ntot
d
1
1
R tot
+ n +
he n =1 n hi

W
2

m K

The thermal transmittance U indicates the rate of heat transfer that flows by a temperature difference of one Kelvin through a surface of one square meter. Thermal insulation is better with low
values of U. The heat flux through the wall can be calculated as follows:
q = U (q i q e )

W
2
m

Steady-State Thermal Transmission

72

i denotes the interior temperature and e the exterior temperature.


The SIA standard 180 [4.2] gives guidelines for the thermal transmittance U of building elements
for heated rooms with regards to thermal comfort and moisture protection (i.e. no surface condensation and mold growth). For energy reasons, however, significantly lower values are almost always required [4.3].

building element

outdoors

unheated room

ground

sloped or flat roof

0.4 W/(m K)

0.5 W/(m K)

0.6 W/(m K)

vertical wall

0.4 W/(m K)

0.6 W/(m K)

0.6 W/(m K)

window, glass door, door

2.4 W/(m K)

2.4 W/(m K)

floor

0.4 W/(m K)

0.6 W/(m K)

0.6 W/(m K)

Tab. 4.1: Maximum thermal transmittance coefficient U for thermal comfort and moisture protection
[4.2]
Figure 4.3 shows the temperature profile for an insulated masonry wall. The temperature jumps at
the wall surfaces can be calculated as follows:

q se q e =

q
he

and q i q si =

q
hi

In layer number n the following temperature gradients occur:

q n
q
=
x n n
d1

d2

dn

i
2

he

se

si
n

hi

1
q

x
Fig. 4.3: Temperature profile in a masonry wall with internal insulation
The coordinate of the location is denoted by x. The temperature in depth xn amounts to:
ntot

( x n ) = e + Rn U ( i e )
n =e

n = e,1,2,..., i

Steady-State Thermal Transmission

73

As mentioned in chapter 1, the temperature profile can also be determined graphically. In Figure
4.4 the calculated and graphically determined temperature profiles in a multi-layered wall are
shown.

Fig. 4.4: Calculated and graphically determined temperature profiles in an external wall [4.4]

74

Steady-State Thermal Transmission

The thermal transmittance from surface to surface of the building element, not including the heat
transfer coefficients he und hi, is denoted by (R = 1/).

Fig. 4.5: Thermal resistances of building materials as a function of the layer thickness [4.5]

Fig.4.6: Thermal resistances of air layers as a function of the thickness [4.5]


The thermal resistance increases proportionally with the thickness of the building material in accordance with the laws of heat conduction for a homogeneous solid. This is graphically presented
in Figure 4.5 for different building materials. In air layers between building materials the heat is
transported by three mechanisms: convection and conduction in the air and radiation exchange between the material surfaces. It should be noted that proportionality of thermal resistance to the
thickness is only valid for heat conduction and not for convection or radiation (Fig. 4.6). Convection

Steady-State Thermal Transmission

75

initiates with a given air layer thickness while the thermal radiation is independent of the air layer
thickness. The high emissivities of many non-metallic surfaces result in the radiant flux becoming
the predominant transport mechanism so that the influence of the air layer thickness in a wider
range becomes minor to the thermal resistance. The radiant flux is of relatively less importance for
metals with low emissivity. After the onset of convection, with an air layer of a couple of centimeters, the thermal resistance decreases after an initial rise following the increasing conduction resistance (for the calculation of the equivalent thermal conductivity of hollow window frames etc. refer to chapter 4.3, and chapter 6.3 for insulated glazings).

4.2

Thermal Bridges

Regions in the building envelope where a higher heat flux occurs than in the standard cross section are designated as thermal bridges. Thermal bridges are relevant with respect to energy losses
as well as respective structural damage. In wintertime, lower temperatures occur on the interior
side at thermal bridges compared with adjacent areas. These locations are therefore critical with
regard to surface condensation and the mold growth. With well-insulated buildings the heat loss
through thermal bridges is no longer negligible as regards heating demand. Therefore, thermal
bridges are to be avoided whenever possible or accordingly their effects should be minimized. This
calls for careful detailing and construction of the building envelope. Figure 4.7 shows possible locations for thermal bridges in a building.

Fig. 4.7: Possible locations for thermal bridges [4.6]


Thermal bridges can be divided into two groups. Material-induced thermal bridges are regions in
the building envelope where a material with a larger thermal conductivity is surrounded by a material with a smaller thermal conductivity. Examples for this are metallic facade anchors, concrete or
steel columns that penetrate the thermal insulation, timber beams between insulation material etc.
Geometry-induced thermal bridges are regions in the building envelope at which a warm area on
the interior side is located across from a larger cold area on the outer side. An external corner of a
building creates such a thermal bridge. Figure 4.8 illustrates these two kinds of thermal bridges.

76

Steady-State Thermal Transmission

Fig. 4.8: Adiabats and isotherms in a material- and a geometry-induced thermal bridge

Fig. 4.9: Adiabats and isotherms in building corners

Fig. 4.10: Temperature distribution on the interior and exterior wall surface of a material-induced
thermal bridge. The one-dimensional (simplified) consideration incorrectly describes the condition
[4.7]

Steady-State Thermal Transmission

77

Figure 4.9 shows adiabats and isotherms in two building corners, in which one is a homogeneous
wall and the other one is a masonry wall with insulation in the middle. The lowest temperature occurs on the interior surface in the corner. Here, conversely, the heat flux is highest. Locations far
away from the corner are handled as a one-dimensional problem (standard cross section).
Figure 4.11 shows the calculated isotherms for a floor slab support. The lowest surface temperature occurs at the inner edge (18C).

Fig. 4.11: Isotherms at a floor slab support. It is assumed that masonry = 0.4 W/(mK) and concrete =
1.8 W/(mK) [4.8]

old

new

Fig. 4.12: Typical building segments old and new [4.6]

Steady-State Thermal Transmission

78

Computer programs have been developed to evaluate the influence of thermal bridges. They are
able to solve two- or three-dimensional steady-state thermal conduction problems (temperature
distributions, heat flows). For planning thermal bridge catalogs are particularly suitable [4.9, 4.10].
In these catalogs typical thermal bridges are shown and their characteristic thermal values listed.
In heat requirement calculations thermal bridges can be taken into account with so-called linear
and point thermal transmittances by means of hand calculations. The total heat flow through a
structural element with a thermal bridge Q tot can be described as the sum of the heat flows
and the additional heat flow due to
through the uninterrupted area (standard cross section) Q
hom

thermal bridges Q Br :
Q tot = Q hom + Q Br

whereby

Q hom = U A ( i e )
Q Br = L ( i e )

or

Q Br = ( i e )

area

m2

length of thermal bridge

thermal transmittance of standard cross section

W/(m2K)

linear thermal transmittance

W/(mK)

point thermal transmittance

W/K

Figure 4.12 shows typical thermal bridges found in older buildings and the analogous construction
in new buildings with a layer of insulation throughout.
In the following, a few strategies to the avoidance of thermally weak locations in the building envelope are expressed. In general:
- conventional construction materials without insulation today produce unsatisfactory results, unless the construction becomes extremely thick
- insulation materials cannot be relied upon to carry loads, with the exception of compressive
loads (foam glass, polystyrene foam), but not for tension or shear
- metallic penetrations can cancel the entire thermal insulation zone of its effectiveness or strongly compromise it
The following guidelines should be adhered to in order to avoid the limitations arising from thermally weak locations:
- consistent wrapping of thermal insulation around the entire building
- no cavities connecting a cold to a warm side of a building element (radiation, convection)
- no metallic penetrations, or at most only punctiform ones (anchors and dowels made out of
plastic when possible)
- supporting structure placed entirely on the warm (interior) side
- access stairways, balconies and similar external loads should be externally supported by pillars
or suspended

Steady-State Thermal Transmission

4.3

79

Cavities in Window Frames etc.

Thermal computations of building elements with unventilated rectangular cavities (window frames,
doors, masonry blocks, etc.) can be determined according to EN ISO 10077-2 [4.11] by an equivalent thermal conductivity:

Fig. 4.12: Rectangular cavity and heat flow direction.

eq =

d
R

d is the length of the cavity in the direction of heat flow and R is the thermal resistance, which is
given by:
R=

1
hc + hr

The convective component of the heat transfer coefficient is given in the case of b > 5 mm by
hc =

C1
d

with C1 = 0.025 W/(mK)

otherwise
1
C

hc = max 1 ; C2 T 3
d

with C1 = 0.025 W/(mK); C2 = 0.73 W/(m2K4/3)

T is the maximum difference of the surface temperature in the cavity.


In the case that the temperature difference is not known, then T = 10 K should be used. This
gives:
C

hc = max 1 ; C3
d

with C1 = 0.025 W/(mK); C3 = 1.57 W/(m2K)

The radiation component of the heat transfer coefficient is:


hr =

F 4 Tm3
1 1
+
1

The view factor F for a rectangular cavity can be calculated as follows:

Steady-State Thermal Transmission

80

F=

2
1
d
d
1+ 1+
2
b
b

s denotes the Stefan-Boltzmann-constant (s = 5.6710-8 W/(m2K4)), 1 and 2 are the emissivities


of each surface and Tm stands for the average temperature in the cavity.
For non-rectangular cavities the cross section can be converted to an equivalent rectangular section (Fig. 4.9) and then, as described above, an equivalent thermal conductance can be calculated.

A = A and

d d'
=
b b'

Fig. 4.9: Conversion of a non-rectangular section into an equivalent rectangular section.

Problems
Problems 1 through 3: Determination of the U-value and temperature profile
The purpose is to become acquainted with the prevalent method to calculate the U-value of building materials and the pertinent means. At the same time one should also develop a sense for the
magnitudes of such values. The problem definition can be found on the following pages.

Problem 4: Facade color and surface temperature


A facade panel with 14 cm insulation ( = 0.04 W/(mK)) is heated on a summer day by the sun (I =
700 W/m2, e = 30C, i = 26C). Determine the exterior surface temperature for a dark ( = 0.8)
and a light ( = 0.2) colored facade. The heat transfer coefficients (infrared-radiation and convection) can be taken as he = 20 W/(m2K) and hi = 8 W/(m2K) respectively. A steady-state condition
can be assumed.

Steady-State Thermal Transmission

Problem 1: Exterior Wall

e = -10C

i = 20C

Wall construction from outside to inside

Material

thickness
(m)

(W/mK)

exterior plaster

0.010

0.87

mineral wool insulating board

0.180

0.036

adhesive mortar

0.010

0.87

masonry block

0.175

0.44

interior plaster

0.010

0.70

81

82

Steady-State Thermal Transmission

Graphic determination of problem 1

Steady-State Thermal Transmission

Problem 2: Roof

e = -10C

i = 20C

roof construction from outside to inside

Material

thickness
(m)

(W/mK)

gravel/extensive plant cover

bitumen

liquid membrane

mineral wool

0.22

vapor retarder
recycled concrete
* negligible thermal resistance

0.036
*

0.31

1.8

83

84

Steady-State Thermal Transmission

Graphic determination of problem 2

Steady-State Thermal Transmission

Problem 3: Floor against cellar

e = 8C

i = 20C

Floor construction from interior to exterior (= cellar)

Material

thickness
(m)

(W/mK)

concrete screed

0.03

1.8

load-distribution plate

0.08

1.5

PE-sheet

impact sound insulation

0.09

0.038

reinforced concrete slab

0.3

1.8

* negligible thermal resistance

85

86

Steady-State Thermal Transmission

Graphic determination of problem 3

Steady-State Thermal Transmission

87

Literature
[4.1]

Zrcher C., Frank T., Bauphysik: Bau und Energie, vdf, Zrich, 2004

[4.2]

SIA Norm 180, Wrme- und Feuchteschutz im Hochbau, Schweizerischer Ingenieur- und
Architekten-Verein, 1999

[4.3]

[4.5]

SIA 380/1, Thermische Energie im Hochbau, Hrsg: Schweizerischer Ingenieur- und Architekten-Verein, 2007
Sagelsdorff, R., Frank, T., element 29: Wrmeschutz und Energie im Hochbau, Schweizerische Ziegelindustrie, Zrich, 1990
Gsele, K., Schle, W., Schall, Wrme, Feuchte, Bauverlag GmbH, Wiesbaden, 1989

[4.6]

Keller B., Bauphysik: Die Energetik des Gebudes, Vorlesungsskript ETH, Zrich, 2006

[4.7]

Cziesielski, E. (Hrsg.), Bauphysik, in: Bautechnik, Bd. V, Konstruktiver Ingenieurbau 2,


Springer-Verlag Berlin, 1988

[4.8]

Brunner, C. U., Nnni, J., Wrmebrckenkatalog 1 (Neubaudetails), SIA D 099, Zrich,


1985

[4.9]

Wrmebrckenkatalog, Bundesamt fr Energie BFE, 2002

[4.4]

[4.10] G. Notter, U.P. Menti, M. Ragonesi, Wrmebrckenkatalog fr Minergie-P-Bauten, In Ergnzung zum Wrmebrckenkatalog des BFE, Schlussbericht, Bundesamt fr Energie
BFE, 2008
[4.11] EN ISO 10077-2, Wrmetechnisches Verhalten von Fenstern, Tren und Abschlssen - Berechnung des Wrmedurchgangskoeffizienten Teil 2: Numerisches Verfahren fr Rahmen, Europisches Komitee fr Normung CEN, Brssel, 2003
[4.12] EN ISO 10211, Wrmebrcken im Hochbau Wrmestrme und Oberflchentemperaturen, CEN, Brssel, 1996
[4.13] EN ISO 14683, Wrmebrcken im Hochbau Lngenbezogener Wrmedurchgangskoeffizient Vereinfachte Verfahren und Anhaltswerte, CEN, Brssel, 1999
[4.14] EN ISO 13370, Wrmetechnisches Verhalten von Gebuden Wrmeverluste ins Erdreich
- Berechnungsverfahren, CEN, Brssel, 1997

Chapter 5
Non-Steady-State Thermal Transmission

Non-Steady-State Thermal Transmission

90

Non-Steady-State Thermal Transmission

In chapter 4 the thermal transmission through a building element was handled with constant temperatures with respect to time on the interior and exterior. In such a steady state the heat flow
through a wall can be calculated by the thermal transmittance U of the building element.
In reality, however, the interior and exterior temperatures are not constant with respect to time but
are subjected to daily and annual fluctuations. The amplitudes of the exterior temperature fluctuations are significantly larger than those on the interior side. Additionally, time-dependent solar radiation can also significantly influence the temperature fluctuations on the exterior surface of the
building envelope. In order to correctly describe the variation of the temperature profile and the
heat flow as a function of time, heat storage in the building element must be accounted for.
In this chapter, therefore, non-steady-state characteristic values shall be developed with which the
reactions of material layers can be described with respect to time-dependent changes of boundary
conditions.

5.1

Non-Steady-State Heat Propagation in a Material Layer

5.1.1 Heat Equation and Illustration


Given is a material layer with a thickness d and thermal conductivity . The surface temperatures
on the interior and exterior are si and se, whereby si > se applies. The condition is steady-state.
The heat flux q then amounts to (Fouriers law of heat conduction):

q q se
q = si
d

(W/m2)

The material layer has a density und a specific heat c. If this layer is now brought from an isothermal state (isothermal = constant temperature throughout the layer) with a temperature 1 into
the isothermal state of temperature 2, whereby 2 > 1, then the quantity of heat stored per m2 in
the layer is given as:
Q = c d ( 2 1 )

(J/m2)

The product of the material quantities c characterizes how well heat can be stored in a material
(unit: J/m3K).
In steady-state conditions the thermal conductivity sufficiently describes the material characteristics. In non-steady-state conditions the additional material quantities c are required. Based on
Fouriers law of heat conduction and the law of conservation of energy a second order partial differential equation can be derived the so-called Fourier differential equation that also describes
the non-steady-state heat propagation (cp. chapter 1.1). In the three-dimensional case x, y and z
denote the coordinates in space and t the time this gives:

2 2 2
=
+
+
t c x 2 y 2 z 2
This differential equation for heat conduction, referred to as the heat equation, describes the temperature distribution in a given region over time. The initial condition, given by the temperature distribution at time t = 0, as well as the boundary conditions are also required in order to solve the
equation in a given case. The boundary conditions can be described by different forms. For example, the temperature or the heat flux can be given at the boundary of the region (note: different
methods to solve Fouriers differential equation analytically or numerically are discussed in the literature, e.g. [5.1]).

Non-Steady-State Thermal Transmission

91

A graphic scheme for the one-dimensional propagation of heat is shown in Fig. 5.1. The material
layer has a constant initial temperature (isothermal). On the left-hand surface the temperature
jumps and is held constant afterwards. The following then takes place:
The temperature increase on the surface causes a heat flow in the first sublayer of the material
( arrow).
Thereby the temperature in the first sublayer is increased. This binds a certain quantity of heat
( bucket).
Because of the temperature increase in the first sublayer a heat flow into the second sublayer is
induced, whereby now less heat flows due to the heat storage in the first sublayer ( slimmer
arrow).
etc.

Fig. 5.1: Model of non-steady-state thermal transmittance (from [5.2])


The material quantities have the following influence:
The larger the thermal conductivity , the better the heat is transmitted.
The larger the storage capacity c, the more the heat is retained and no longer available to
propagate.
5.1.2 Thermal Diffusivity and Thermal Effusivity
A more exact treatment shows that with the two quantities and c as a product or as a ratio two
new important characteristic values can be generated, which describe heat propagation.
The term /(c) is denoted as thermal diffusivity a and describes the speed and range of the temperature equalization in a material layer. The range is larger, the larger the thermal conductivity
is (ability to transfer) and the range is smaller, the more is retained in transit c (fill and empty
buckets).

Non-Steady-State Thermal Transmission

92

Thermal diffusivity:

a=

m2

The thermal effusivity b is a measure of how much heat is lost by propagation in a material. This
will be stronger the greater the heat conductivity is (long-range capability) and the greater the
storage capability c is (larger heat storage).
Thermal effusivity:

2
1/2
m K s

b = c

In Table 5.2 the properties , and c for different materials are shown as well as thermal diffusivity
a and thermal effusivity b.

Material
copper
aluminum
steel
concrete
reinforced-concrete
brick (clay)
sand lime brick
earth (sand/gravel)
earth (clay)
float glass
gypsum plasterboard
wood (spruce)
chipboard
mineral wool
expanded polystyrene
air (still, 20C)
water (still, 10C)

(W/mK)
380
200
50
1.65
2.3
0.44
1.0
2.0
1.5
1.0
0.25
0.14
0.14
0.04
0.04
0.026
0.6

(kg/m )
8900
2700
7800
2200
2300
1100
1800
2000
1500
2500
900
500
600
80
15
1.2
1000

c
(J/kgK)
380
900
450
1000
1000
900
900
1000
2000
750
1000
2200
1700
600
1400
1000
4190

c
3
(kJ/m K)
3'382
2'430
3'510
2200
2'300
990
1'620
2'000
3'000
1'875
900
1'100
1'020
48
21
1.2
4'190

a
b
2
2
1/2
(m /s) (kJ/m Ks )
-8
1123610
35.8
-8
823010
22.0
-8
142410
13.2
-8
7510
1.9
-8
10010
2.3
-8
4410
0.7
-8
6210
1.3
-8
10010
2.0
-8
5010
2.1
-8
5310
1.4
-8
2810
0.5
-8
1310
0.4
-8
1410
0.4
-8
8310
0.04
-8
19110
0.03
-8
216710
0.006
-8
1410
1.6

Tab. 5.2: Characteristic values of different materials (from [5.3, 5.4])


If two semi-infinite bodies with different temperatures 1 und 2 come into contact, the temperature
0 at the contact surface can be calculated by means of the thermal effusivities b1 and b2 :

0 =

b1 1 + b2 2
b1 + b2

(C or K)

Lets assume that a persons body (body = 34C, bbody = 1 kJ/m2Ks0.5) comes into contact with a
thick plate that is at room temperature. The plate is either made of polystyrene (bPS = 0.03
kJ/m2Ks0.5) or copper (bCu = 35.8 kJ/m2Ks0.5). The following applies: bCu > bbody > bPS. The temperature of the material with the larger thermal effusivity (better dissipation of heat) has a higher
impact on the contact temperature:

0,PS =

1000 34 + 30 20
= 33.6C
1000 + 30

0,Cu =

1000 34 + 35800 20
= 20.4C
1000 + 35800

Since the contact temperature with the copper plate is considerably lower, ones touch feels colder
than with the polystyrene plate.

Non-Steady-State Thermal Transmission

93

To illustrate the impact of thermal diffusivity, the temperature development in a 40 cm thick wall is
shown in Figure 5.2 after a sudden temperature change on the surface (0C 20C). The backside of the wall is adiabatic (= perfectly insulated, i.e. no heat flows through the backside surface). The rate of temperature development differs according to each material: wood, concrete or
polystyrene. It gives: aH < aB < aPS.

Temperature (C)

Wood
aH = 1310 m /s
-8

20

48

10
36
24

0
0

12
20

Wall thickness

Time (h)

40 0

Temperature (C)

Concrete
aB = 7510 m /s
-8

20

48

10
36
24

0
0

Time (h)

12
20

Wall thickness

40 0

Temperature (C)

Polystyrene
aPS = 19110 m /s
-8

20

48

10
36
24

0
0

12
20

Wall thickness

40

Time (h)

Fig. 5.2: Temperature development after a surface temperature jump [5.2]


5.1.3 Inertia of Walls: Charge and Discharge Performance
Often the boundary conditions of a wall suddenly change; outside, e.g., with a cold spell or a storm,
inside, e.g., if a window is opened or a heater is turned on or off. It is, therefore, interesting to know
how quickly a wall reacts to a sudden change of the ambient temperature. That is, one would like
to be able to assess the thermal inertia of the wall. The transition from one steady-state condition
to another one is named transient.
As an example, Figure 5.3 shows the discharge (cooling down) of a 20 cm thick concrete wall.
The wall and its surroundings are for t < 0 at a constant temperature i. At time t = 0 the ambient
temperature abruptly sinks. The change in the temperature profile in the wall up to t = 36 h as well
as the heat flow (normalized) to the surroundings are shown in Figure 5.3.

Non-Steady-State Thermal Transmission

94

d = 0.2 m
1

t=0

t=2h

t=6h
t=8h
t = 10 h
t = 12 h

Heat flow

t=4h

0.8

0.4
0.2

t = 24 h

0.6

t = 36 h

10

20

30

Time (h)

Fig. 5.3: Discharge (cooling down) of a concrete wall after a temperature jump (i f)
The cooling proceeds more slowly, the larger the layer thickness d is and the smaller the thermal
diffusivity a is. With these two quantities a characteristic time d2/a can be computed. During the
cooling process the surface temperature decreases approximately exponentially and the characteristic time d2/a can very roughly be interpreted as a time constant [5.6]:

d2
a

(s)
1
t

With an exponential decay in time (e.g. (t ) = (t = 0 ) e t ) the time constant denotes the time
with which the value subsides to the fraction 1/e (Eulers number e = 2.71828, thus 1/e = 36.8%).
The heat transfer coefficient to the surroundings likewise influences the cooling action. This is here
however negligible. Time constants for a couple of materials with a thickness of 10 cm are:
Wood:

0.1 2
1

21h
8
3600
13 10

Concrete:

0 .1 2
1

4h
8
3600
75 10

Mineral wool:

1
0 .1 2
3h

8
3600
83 10

Copper:

1
0 .1 2

1.5 min
8
3600
11'236 10

The term d2/a can be expressed as follows:


d2 d2 c d
=
= c d = R C

That is, the time constant is larger, the higher the thermal resistance R and the larger the surface
storage capacity C of the layer are. Lets consider two material layers.
Mineral wool (d = 30 cm):

Non-Steady-State Thermal Transmission

R C =

c d =

95

0 .3
80 600 0.3 = 7.5 14 '400 = 30 h = 1.25 d
0.04

Masonry made of lightweight bricks (d = 47.5 cm, = 0.17 W/mK, = 960 kg/m3, c = 900 J/kgK):

R C =

c d =

0.475
960 900 0.475 = 2.79 410 '400 318 h 13 d
0.17

The layer of mineral wool insulates (despite less thickness) almost 3 times better than the masonry
made of lightweight bricks, however it has almost 30 times less storage capacity. As a result the
masonry has more thermal inertia, by about a factor of 10!
After a period of about three time constants (e-3 0.05), the previous state is almost completely
forgotten and the system is in steady state. This occurs with the mineral wool layer after about 4
days, the masonry however needs noticeably longer than a month for this! Since the boundary
conditions during this time have for the most part already changed, such walls are practically never
in a complete steady state.
5.2

Reaction of a Material Layer to Periodic Changes

5.2.1 Material Layer with a Finite Thickness: Amplitude Damping und Phase Shift
Very often the boundary conditions fluctuate periodically. Daily and annual variations in the outside
air temperature resemble a sine or cosine function (cp. chapter 2). It is, therefore, helpful to understand how a material layer reacts to a harmonic variation of the boundary conditions.
The surface of a material layer (x = 0) is subjected to a harmonic temperature oscillation with an
amplitude at an average temperature 0:
2
t
T

(t , x = 0 ) = 0 + cos

(C)

The backside of the material layer is adiabatic. The temperature oscillation propagates from the
front surface into the material as a heat wave. In doing so the temperature amplitude reduces
with increasing depth.
Amplitude damping: =

'

(-)

With increasing depth into the material the temperature oscillation also sustains a lag. That is, the
maximum and the minimum temperature shift back in time (Figures 5.4 and 5.5).
Phase shift:

(s or h)

For the temperature oscillation at depth x, it gives:

2
(t ')

(t , x ) = 0 + ' cos

(C)

Non-Steady-State Thermal Transmission

96

48

Temperature (C)

20

42
36
30

Time t (h)

24

10
18
12
0

6
0

10

20

Phase shift

30

Material depth x

Fig. 5.4: Temperature development in a concrete wall with periodic excitation [5.2]
Phase shift e

15

Amplitude ''

Amplitude

Temperature (C)

20

10

12

24
Time (h)

36

48

Fig. 5.5: Front and back side surface temperature with periodic excitation

5.2.2 Semi-Infinite Material Layers: Penetration Depth


In the case of the surface of a semi-infinite material thickness subjected to a harmonic temperature
oscillation, the time dependent temperature profile can be calculated simply. This case corresponds, for example, to the penetration of a temperature oscillation into the ground (chapter 2) or
into a very thick wall. The temperature oscillation in the semi-infinite layer is given by:
x
2
t

(t , x ) = 0 + e x / co

(C)

Non-Steady-State Thermal Transmission

T
T
=
a
c

Penetration depth:

97

(m)

With increasing depth the temperature amplitude decreases (Fig. 5.6). The penetration depth s
denotes the depth in which the amplitude is reduced to the 1/e-part (1/e = 36.8%, Fig. 5.7). The
penetration depth characterizes the range of temperature oscillation and is larger the larger the
thermal diffusivity a of the material is, but also the larger the period T of the oscillation is.
1
Concrete a = 75 .10 m /s
Period T = 24 h
Penetration depth s = 0.144 m
-8

t=0
t=3h

Temperature (K)

0.5

t = 21 h

t=6h
t = 18 h

t=9h

-0.5

t = 15 h

t = 12 h

-1

0.1

0.2

0.3

0.4

0.5

Depth x (m)

Fig. 5.6: Penetration of a temperature oscillation into a semi-infinite material layer


100 %

36.8 %
13.5 %
5.0 %
2

2 e

x=s

2 e

x = 2s

2 e

x = 3s

Fig. 5.7: Envelope of temperature oscillations in a semi-infinite material layer and penetration
depth s
The penetration depth s is proportional to
layer about

T . Thus the annual oscillations penetrate the material

365 19 -times deeper than the daily oscillations. Table 5.3 shows the dependence

Non-Steady-State Thermal Transmission

98

of the penetration depth of different materials on the period. The annual oscillations in particular
show penetration depths of more than a meter. These penetration depths are considerably larger
than commonly constructed layer thicknesses. Slow oscillations go right through the building construction, as if the building was transparent! Faster oscillations from about T = 24 h to T = 1 h can
still enter common layer thicknesses. Very fast oscillations have an effect only on the near surface.

Material
aluminum
steel
concrete
reinforced concrete
brick (clay)
sand lime brick
earth (sand/gravel)
earth (clay)
gypsum plasterboard
wood (spruce)
mineral wool
expanded polystyrene

1h
0.307
0.128
0.029
0.034
0.023
0.027
0.034
0.024
0.018
0.012
0.031
0.047

4h
0.614
0.256
0.059
0.068
0.045
0.053
0.068
0.048
0.036
0.024
0.062
0.093

12 h
1.064
0.443
0.102
0.117
0.078
0.092
0.117
0.083
0.062
0.042
0.107
0.162

Period T
24 h
1.505
0.626
0.144
0.166
0.111
0.130
0.166
0.117
0.087
0.059
0.151
0.229

1 Week
3.981
1.656
0.380
0.439
0.293
0.345
0.439
0.310
0.231
0.157
0.401
0.606

1 Month
8.241
3.428
0.787
0.908
0.606
0.714
0.908
0.642
0.479
0.324
0.829
1.254

1 Year
28.744
11.958
2.744
3.168
2.112
2.489
3.168
2.240
1.670
1.130
2.892
4.373

Tab. 5.3: Penetration depths s (m) for different periods and materials
With non-steady-state processes the penetration depth can be used as a measurement for the
thermal effective layer thickness in a particular material. d denotes the geometric thickness of the
material layer.
Dynamic layer thickness:

z=

(-)

Because the penetration depth depends on the period of excitation, the dynamic layer thickness
also depends on the period. A wall with a thickness d can also appear thin (s large z small) for
slow oscillations, however thick (s small z large) for fast oscillations.
The quantity of energy that flows into a semi-infinite material layer per period, that is stored there
and subsequently flows out again, is given by:
QT = 2

T
T
c = 2
b
2
2

(J/m2)

denotes the amplitude of the temperature oscillation on the surface. The thermal effusivity b is a
measurement of how much heat penetrates into a material layer. Since the stored heat is proportional to T , with annual oscillations more heat is stored than with daily oscillations.
Lets consider as an illustration the following example. For an assessment we calculate the penetration depth of a wall. If good and bad weather periods follow one another, the outside temperature rises and falls accordingly, so that the daily oscillations of a slower fluctuation becomes superimposed (Fig. 5.8). The course of the outside temperature can then be depicted as the superposition of two harmonic oscillations:

Non-Steady-State Thermal Transmission

99

External temperature (C)

30
25
20
15
10

May 22 - June 1, 2005


Zurich SMA

5
1

10

11

Time (h)

Fig. 5.8: Course of the instantaneous and daily average temperature


2
2
t + 2 cos
t
T1

T2

(t ) = 0 + 1 cos

(C)

Assuming the temperature amplitudes and periods according to Figure 5.8 and a light exterior wall
construction with 30 cm of mineral wool insulation (U = 0.13 W/m2K) it is obtained for penetration
depths and dynamic layer thicknesses:
Daily oscillation:

1 = 7 K, T1 = 24 h

s1 = 0.151 m

z1 =

0 .3
= 1.99
0.151

e z1 = e 1.99 = 0.137

10-day-oscillation:
0 .3
e z1 = e 0.627 = 0.534
= 0.627
0.479
The daily oscillation virtually does not penetrate through the wall construction (13.7%), the 10-day
oscillation, however, becomes significantly less damped (53.4%).

2 = 6 K, T2 = 240 h

s2 = 0. 479 m

z2 =

5.2.3 Effective Thickness for Heat Storage


The capability of the interior building elements to store heat and then to release it again later in
time often plays a significant role for the non-steady-state behavior of a room or building (chapter
8). If walls, ceilings and floors have a large heat capacity, the temperature fluctuations can be
smoothed out. This is advantageous with regards to thermal comfort (chapter 3).
In section 5.2.2 it was shown that the layer thickness in which heat is stored varies depending on
material characteristics and frequency of excitation. This effective storage layer thickness can be
characterized by the penetration depth s. For a material layer with a thickness d the dynamic heat
capacity per area Cdyn can be calculated as follows:
For thin material layers with d
For thick material layers with d >

1
2

1
2

Cdyn = c d

: Cdyn = c

(J/m2K)

1
2

(J/m2K)

Non-Steady-State Thermal Transmission

100

With a thin layer the total thickness d is active in heat storage while with a thick layer only the
outer part is. The division between thin and thick is based on the penetration depth; i.e. on the
material properties and the period of excitation. It follows by means of the thermal effusivity b:
Cdyn = c

= c

1
2

T
T
T
=
c =
b
c
2
2

That is, with a thick layer the storage capacity does not dependent on the thickness d! Figure 5.9
shows the dynamic heat capacity with respect to the layer thickness, plotted as multiples of the
penetration depth s .

Dynamic heat capacity ratio


Cdyn/Cdyn, (-)

thin

thick
Cdyn,

1.0
0.8

d=

0.6

s
2

0.71 s

0.4
0.2
0

2s

3s

4s

5s

Layer thickness

Fig. 5.9: Dynamic heat capacity ratio of a material layer as a function of multiples of the penetration
depth s
For daily oscillations (T = 24 h) the effective storage layer thickness for typical massive construction materials is only about 10 cm. If the wall is thicker, then with daily oscillations the additional
layer thickness is essentially under-utilized (Tab. 5.4). Also an infinitely thick wall possesses only a
finite dynamic storage capacity!

Penetration depth s
for T = 24 h
(m)

Effective storage layer


thickness / 2

Dynamic
heat capacity Cdyn

(m)

(kJ/m2K)

concrete

0.144

0.102

224

sand lime brick

0.130

0.092

149

brick (clay)

0.111

0.078

78

Tab. 5.4: Heat storage characteristics of material layers with daily fluctuations

5.3

Non-Steady-State Properties of Opaque External Walls

Lets consider an opaque wall in which the exterior air temperature is oscillating harmonically. In
section 5.2.1 it was shown that with increasing depth of wall the temperature oscillation experiences a phase shift and a dampening of its amplitude. The amplitude of the temperature oscillation on
the interior wall surface is thus smaller than that of the exterior temperature and the fluctuation exhibits a shift in time of the maxima (phase shift ). The ratio of the amplitudes of the exterior air

Non-Steady-State Thermal Transmission

101

temperature to the amplitude of the interior surface temperature is denoted as amplitude damping
(Fig. 5.10).

si (t)
e (t)

si (t)

si

e
si

Fig. 5.10: Amplitude damping and phase shift


The heat transmission for a periodic excitation and multi-layered walls can also be presented by
Heindls approach in a matrix [5.7, 5.8, 5.9]:
q e W11 W12 q i

q e W21 W22 q i

The building element properties are contained in the heat transfer matrix W, which describes the
relationship between the temperature and the heat flow oscillations inside and outside. The matrix
elements are represented by complex numbers. A more exact consideration shows that three
properties characterize the non-steady state:
amplitude damping
phase shift
dynamic thermal transmittance UT
In order to determine these characteristic values, the internal boundary conditions must be defined.
In doing so two boundary cases are differentiated (Fig. 5.11).
The boundary condition I (i = constant) approximates the case of an air-conditioned building or a
building with a large thermal inertia. The boundary condition II (qi = 0) corresponds rather to the
case of a building with little thermal inertia.
Because the heat equation is a linear equation the different solutions can be superimposed. If for
boundary condition I the solution for the steady-state case is superimposed onto the solution for
the non-steady-state case, the extreme heat fluxes can be expressed as follows:
max
max
q i min = q stat q instat min = U (q i q e ) UT q e

(W/m2)

internal air temperature

mean value of the external air temperature

amplitude of the external air temperature

U
UT

thermal transmittance (steady-state)


dynamic thermal transmittance (non-steady-state portion)

W/m2K
W/m2K

Non-Steady-State Thermal Transmission

102

The dynamic thermal transmittance UT corresponds to the maximum heat flux on the internal surface of the wall if the external air temperature oscillates periodically by 1 K.
Characteristic values:

Boundary condition I: Isothermal


Exterior

he

hi

Interior

- amplitude damping T

.
qi 0

- phase shift T

i = constant

- dynamic resistance RT =
- dynamic U-value UT =

- reduction factor fT =

e
.
q i

1
RT

UT
U

Boundary condition II: Adiabatic


Exterior

he

Interior

Characteristic values:

qi = 0 (adiabatic)

- amplitude damping H
- phase shift H

Figure 5.11: Boundary conditions to calculate the non-steady-state characteristic values


Figure 5.12 shows the non-steady-state characteristic values of differently constructed external
walls. These properties apply to the idealized boundary conditions in accordance with Figure 5.11.
It can be seen from figures 5.12 und 5.13 that the layer sequence influences the non-steady-state
characteristic values. To dampen the external temperature oscillation as much as possible, a wall
with external insulation is better suited than a wall with internal insulation. The better a wall is insulated and the more mass per unit area is available, the higher the amplitude damping and the larger the phase shift. High values of amplitude damping and phase shift are beneficial with respect to
thermal comfort. Poorly insulated light wall constructions exhibit limited amplitude damping and
phase shifting!
The idealized boundary conditions described are in reality practically never existing, so that the
non-steady-state parameters , und UT are only suitable for rough estimates.
By way of illustration lets compare the fluctuation of the heat fluxes of two different wall constructions with identical thermal transmittances (U = 0.27 W/m2K) on a nice summer day ( e = 24C ,

e = 10 K , i = 22C) and a winter day ( e = 5C , e = 5 K , i = 20C). A negative sign corresponds to an inward heat flow:
(i)

Externally insulated massive wall (1 cm external rendering / 12 cm insulation / 25 cm concrete


/ 1.5 cm internal plaster)
Summer:

max
max
q i min = q stat q instat min = 0.27 (22 24 ) 0.03 10 = 0.54 0.3W / m 2

Winter:

max
max
q i min = q stat q instat min = 0.27 (20 ( 5 )) 0.03 5 = 6.75 0.15W / m 2

Non-Steady-State Thermal Transmission

103

Randbedingung I:
Isotherm
(i = 0)

UT

Randbedingung II:
Adiabatisch
(qi = 0)

Fig. 5.12: Non-steady-state characteristic values of external wall constructions [5.5]

Non-Steady-State Thermal Transmission

104

(ii) Lightweight wall construction (0.1 cm aluminum / 14 cm insulation / 0.1 cm steel)


Summer:

max
max
q i min = q stat q instat min = 0.27 (22 24 ) 0.26 10 = 0.54 2.6W / m 2

Winter:

max
max
q i min = q stat q instat min = 0.27 (20 ( 5 )) 0.26 5 = 6.75 1.3W / m 2

The fluctuation portion is with lightweight construction considerably more significant than with massive construction and can easily surpass the steady-state portion in summer!
External insulation
Insulation
0.2 m

Concrete
0.2 m

Amplitude damping
H = 103

Internal insulation
Concrete
0.2 m

Insulation
0.2 m

Amplitude damping
H = 3.6

Fig. 5.13: An externally insulated wall dampens an external temperature oscillation better than a
wall insulated on the interior (T = 24 h; boundary conditions: he = 25 W/m2K, interior adiabatic).

5.4

Structural Consequences

In order to ensure high thermal comfort in a building, the temperature fluctuations in the interior
rooms must be kept small. This should be achieved, whenever possible, without HVAC (heating,
ventilation and air conditioning) systems. Therefore, an appropriate building construction is essential.
Massive construction that is well insulated on the exterior is most suitable in keeping the effect of
the external temperature oscillations on the interior temperature small. Such building construction
exhibits high amplitude damping whereby the impact of the phase shift is negligible since the amplitudes are only tiny (Note: In the past one was worried that the effect of the external oscillations,
for example, only reach the interior after working hours).
Because of absorption of solar radiation on the external surface of walls and roofs the fluctuations
of the external surface temperature can be significantly raised. Light facade and roof surfaces, roof
areas possibly with green vegetation, and possibly a ventilated air space reduce the effect of solar
radiation (cp. Fig. 2.25).
The heat storage capacity of the interior structural elements (walls, ceilings and floors) is important
to keep the effects from fluctuations of released heat solar radiation, heat from people and appliances on the interior temperature small. Massive construction made of concrete or sand-lime
bricks, for example, are especially good for this. However, it is important that the thermal resistance between the indoor air and the building element is not increased by suspended ceilings,
raised floors, thick carpets etc.

Non-Steady-State Thermal Transmission

105

Problems
Problem 1: Heat storage in a building element
Given are three different material layers:
Thermal conductivity
W/(mK)

Density
3

kg/m

Specific heat

Thickness

J/(kgK)

concrete (reinforced)

2.3

2300

1000

0.2

gypsum plasterboard

0.25

900

1000

0.025

mineral wool

0.04

60

1080

0.2

Determine each of the following for a daily temperature oscillation as well as a monthly oscillation:
a.) penetration depth
b.) characterization of the layer with respect to the dynamic thermal properties (thick or thin)
c.) dynamic heat capacity
Compile your results in a table; one for the daily oscillations and one for the monthly.

Problem 2: Underground exhibition hall


In the immediate vicinity of the pyramids in Gizeh (Egypt), an underground exhibition hall for visitors is planned. The distance between the ground surface and the exhibition hall is a minimum of 5
m. The annual mean temperature in Gizeh is 21C, the annual amplitude is 10 K and the temperature diffusivity of the ground can be taken as 10-6 m2/s.
a.) How large is the penetration depth of the daily and annual oscillations?
b.) Mathematically show in which range the temperature fluctuates at a depth of 5 m during a
year?
c.) How great is the amplitude damping for an annual oscillation at a depth of 5 m?
d.) How does the room temperature change when there are visitors in the room? Give reasons to
support your answer qualitatively.
Problem 3: Time constant of walls
Given are two walls: a sand lime brick wall (d = 20 cm) and a wall constructed of spruce wood (d =
2 cm). Evaluate mathematically:
a.) time constant of the sand lime brick wall
b.) time constant of the wall constructed of spruce wood
Problem 4: Contact temperature of floors
Three people stand barefoot each for a long period of time on a floor of wood, stone and steel.
The room temperature is at 20C, the temperature of the people can be taken as 34C. The thermal effusivity for a persons body is bbody = 1 kJ/m2Ks0.5 and for the floors is
bwood = 0.4 kJ/m2Ks0.5, bstone = 1.5 kJ/m2Ks0.5 and bsteel = 13.2 kJ/m2Ks0.5.
a.) Calculate the contact temperature for all three floors.
b.) Interpret your results from a.).

Non-Steady-State Thermal Transmission

106

Literature
[5.1]
[5.2]
[5.3]
[5.4]

Grigull U., Sandner H., Wrmeleitung, Springer-Verlag, Berlin, 1990


Keller B., Bauphysik: Die Energetik des Gebudes, Vorlesungsskript ETH, Zrich, 2006
SIA 279, Wrmedmmstoffe, Schweizerischer Ingenieur- und Architekten-Verein, Zrich,
2004
SN EN 12524, Baustoffe und -produkte - Wrme- und feuchteschutztechnische Eigenschaften - Tabellierte Bemessungswerte, Europisches Komitee fr Normung, Brssel, 2000

[5.5]

Zrcher C., Frank T., Bauphysik: Bau und Energie, vdf, Zrich, 2004

[5.6]

Keller B., Rutz S., Pinpoint: Fakten der Bauphysik zu nachhaltigem Bauen, vdf Hochschulverlag, Zrich, 2007

[5.7]

Sagelsdorff R., Frank T., element 29 - Wrmeschutz und Energie im Hochbau, Schweizerische Ziegelindustrie, Zrich, 1990

[5.8]

SIA 180, Wrme- und Feuchteschutz im Hochbau, Schweizerischer Ingenieur- und Architekten-Verein, Zrich, 1999

[5.9]

EN ISO 13786, Thermal performance of building components - Dynamic thermal characteristics - Calculation methods, Europisches Komitee fr Normung, Brssel, 199

Chapter 6
Transparent Building Elements

108

Transparent Building Elements

Transparent Building Elements

Transparent elements of the building facade let daylight into the building and allow for a view from
the inside to the outside. They also let in incident solar energy for heating the building in winter. To
save on heating, the thermal losses in winter through the transparent building elements should be
small. In general, thermal transmittances of glazings are significantly higher than those of adjacent
opaque elements. To reduce heating demands it is desirable that glazings have a small thermal
transmittance and likewise a high total solar energy transmittance. A small thermal transmittance is
also advantageous regarding thermal comfort because the internal surface temperature of the
glazing will then be close to the interior air temperature (chapter 3). In order to avoid the building
overheating in summer, solar shading is mostly necessary. Thereby, the amount of solar energy
that comes into the building interior can be controlled. The solar shading device can significantly
influence the facade design and should, therefore, be included early on in the design and construction process. Especially in work places attention should also be given to suitable measures to
avoid the effects of glare.
In this chapter the characteristics of transparent building elements shall be introduced and the
transport of solar radiation and heat through glazings and windows shall be investigated.

6.1

Classification and Characteristics

Basically there are several ways to fabricate transparent insulated building elements. Figure 6.1
shows a geometrical classification. The designation transparent is for some of these materials not
actually applicable. The term translucent would be more suitable to apply because the linear propagation of light in some of these structures is hindered by scattering or reflection and it is not possible to look through them (objects on the other side can not be clearly discerned). However given
that the term transparent has already been established, it will be used here as well.

Glass panes, possibly with films

Honeycomb or capillary structures

Foams, closed-cell structures

Microporous materials (aerogel)

Fig. 6.1: Classification of transparent building elements [6.1]


For each of the four categories in Figure 6.1, examples of technical execution are illustrated. With
glazings the individual layers panes of glass or possibly plastic films are arranged parallel to
one another. If honeycomb or capillary structures are placed between two cover sheets (e.g. out of
glass), then the solar radiation will be reflected forwards in the direction of the absorber or interior
room. Thereby high transmittances of solar irradiance can be achieved. Large pore structures or

Transparent Building Elements

109

cavity structures are e.g. foams or a combination of layers that lay parallel and perpendicular to the
bounded face sheet. When the pore size of microporous materials is significantly smaller than the
wavelength of the sunlight, reflection will no longer occur in the pores. Thereby the material appears homogeneous and is transparent.
Today in buildings, glazings are almost exclusively used as transparent building elements. Only
with glazings is it possible to clearly see through and moreover the thermal as well as the optical
properties have significantly improved in recent years so that today they can be referred to as
technically superior.
If the solar irradiance strikes a transparent building element, part of the incident energy will be reflected (e), i.e. directed back to the exterior, a part will be absorbed (e) and a part will be transmitted (e) from the building element into the building interior (Fig. 6.2). Based on the conservation
of energy one obtains:

e + e + e = 1

(-)

The absorbed solar irradiance produces a warming of the building element. A part of this heat
flows into the building interior. The total solar energy transmittance g denotes the proportion of the
incident solar energy that is converted into heat in the interior and consists of the solar transmittance e and the secondary internal heat transfer factor qi.
Total solar energy transmittance:

g = e + qi

(-)

The solar transmittance e depends on the incidence angle of solar radiation, the material properties and on the geometric structure of the transparent building element. Also the secondary internal
heat transfer factor qi depends on these quantities and is ultimately determined by the surface
temperature occurring on the interior side as well as the heat transfer coefficient.
Transparent building element
Incident solar radiation
Solar transmittance e
Solar reflectance e
Solar absorption e

Total solar energy


transmittance
g = e + qi

Secondary internal
heat transfer factor qi

Fig. 6.2: Definition of solar transmittance e, secondary internal heat transfer factor qi and total
solar energy transmittance g
The thermal transmittance U results from the heat transport mechanisms in the transparent building element conductance in the material, conductance and convection in the gas filling and
thermal radiation plus the heat transfer coefficients at the exterior and interior surfaces.
Figure 6.3 shows energy properties of different glazings. Plotted are the total solar energy transmittance g and the thermal transmittance U. In order to minimize the heating demand in winter, glazings with U-values (losses) as small as possible and g-values (gain) as large as possible are most
suitable. A data point in the upper left corner of the figure corresponds to the highest gain/loss ratio. The more layers of glass panes and gaps in a glazing, the higher will be its thermal resistance
( smaller U-value), but the solar radiation that passes through will be less ( smaller values of e

Transparent Building Elements

110

Total solar energy transmittance g (-)

and g). The optical and thermal processes that determine these specific values will be investigated
in the following.

1
0.8
0.6

Double glazings
0.4

Triple glazings
0.2
0
0

0.5

1.5

2.5

Thermal transmittance U (W/(m K))

Fig. 6.3: Total solar energy transmittance and thermal transmittance for double and triple glazings

6.2

Optical Properties of Glazings

In this section it will be examined how solar radiation is transmitted through a glazing (see e.g. also
[6.2]). Given is a planar transparent layer that is absorbing but not scattering. The conservation of
energy requires for a ray of light:
A+R+T=1
A designates the absorption, R the reflection and T stands for the transmission. Fresnels Law for
reflection of unpolarized light on a boundary layer between two mediums 1 and 2 gives (Fig. 6.4):
R=

1 sin2 (2 1 ) tan2 (2 1 )

2 sin2 (2 + 1 ) tan2 (2 + 1 )

The angle denotes the deflection of the light rays from a normal to the boundary layer. According
to Snellius Law, n1 and n2 denote the refraction indices of both mediums, it gives:
n1 s i n1 = n2 s i n2

For normal (1 = 0) incident light it gives:


n n2
R = 1
n1 + n 2

For a glass-air-boundary layer with n1 1 (air) and n2 1.5 (glass) and for normal incident light, the
result is a degree of reflection R = 0.04.

Transparent Building Elements

111

1
R

n1 < n 2
R+T=1
n1
n2

Fig. 6.4: Reflection und refraction of a light ray on a boundary layer between two mediums with
different refraction indices
The intensity of a light ray decreases exponentially as it traverses a medium. According to the
Bouguer-Lamberts Law the transmission factor T' for a light ray in a medium with an absorption
coefficient and the local coordinate x in the direction of the light ray is:
T ' = e x

Figure 6.5 shows how a light ray strikes a transparent layer (e.g. a glass pane). The amount of the
total reflectance, absorptance and transmittance, , and , can be calculated from the quantities
R and T', that are given by the material properties and the angle of incidence.

R(1-R) T
2

air n1 1
glass n2 1.5
T
air n1 1
(1-R) T (1-R) R T
2

Fig. 6.5: Light paths in traversing a glass pane


The part R of the striking light is reflected. The part (1 R) penetrates the material and is weakened by crossing through the layer due to absorption by the factor T'. The total transmittance is a
result of the summation of the individual transmitted rays:

= (1 R )2 T ' 1 + R 2T '2 +R 4T '4 +...

This infinite geometric series can be added as follows:

= ( 1 R ) 2 T ' R 2 mT ' 2 m =
m =0

( 1 R )2T '
1 R 2T ' 2

With the same approach one obtains the total reflectance :

= R 1 +

2
T '2 (1 R )

1 R 2T '2

Transparent Building Elements

112

Based on the conservation of energy the total absorptance can now also be calculated:

=1--
With very weak absorbing glazings is T' 1 and the total reflectance is determined only by the refraction index:

2R
1+ R

absorptance (%)

Transmittance, reflectance,

For a glass pane surrounded by air, very weak absorbing glass panes and normal light incidence
(R = 0.04) one obtains = 0.077. According to Fresnels Law the reflectance increases with increasing angle of incidence. With the same procedure of tracing the path of the light ray as shown
in Figure 6.5, the values , 1, 2 and of a double pane glazing can also be deduced. Figure 6.6
shows for a double glazing the angle dependence of reflectance and transmittance , as well as
the absorptance in the two glass panes 1 and 2. For 1 90 all radiation will be reflected.

Incidence angle ()
Fig. 6.6: Influence of the incidence angle of the solar radiation on transmittance, reflectance and
absorptances (external and internal pane) of a double glazing [6.3]
The optical properties of transparent layers are in general dependent on the wavelength. Figure
6.7 shows the influence of the iron-oxide content on the spectral transmittance of a glass pane.
Iron-oxide is an impurity that gives the glass a green tint. To reach high solar transmittances, e.g.
in solar collectors, glass panes with limited iron-oxide content are used. Glass is opaque to radiation with wavelengths of > 3 m. For thermal radiation of bodies at room temperature, glass is,
therefore, opaque.

Transparent Building Elements

113

Transmittance ()

1.0
0.02% Fe2O3
(colorless glass)
0.10% Fe2O3
(standard-float-glass)
0.5

0.50% Fe2O3
(green glass)
0
0.2

1.0

2.0

3.0

Wavelength (m)

Fig. 6.7: Influence of iron-oxide content on the spectral transmittance of 6 mm thick glass panes
with normal light incidence [6.4]
The integral transmittance of a glazing for solar irradiance I(,) is dependent on both the wavelength and the angle of incidence. For a specific incident angle the solar transmittance can be
calculated as follows:

e ( ) =

( , ) I ( , ) d

(-)

I( , ) d
0

The integral reflection can be determined analogously.


In the determination of the transmittance of visible light v, the eyes sensitivity will additionally be
considered [6.6]. The light transmittance with regard to providing daylight into the interior is also an
important characteristic value of the glazing (see Building Physics III).
By means of very thin coatings (ca. 0.1 m), notably metallic, the optical and thermal properties of
the glazing can be modified. Insulating glazings should have a high transmittance for solar radiance (0.3 m 3 m), including light (0.38 m 0.78 m) and a little thermal transmittance.
To avoid a possible cooling load in a building it can be worthwhile to employ a glazing that selectively transmits solar radiation. Such solar control glazings should have a high light transmittance
(0.38 m 0.78 m), yet only a small transmittance for infrared solar radiance (0.78 m 3 m).
Figure 6.8 shows the spectral transmittance and reflection of a solar control glazing.
The spectral selectivity indicates how selective solar radiation is transmitted corresponding to the
ratio between light transmittance v and the total solar energy transmittance g.
Spectral Selectivity:

S=

v
g

(-)

Solar control glazing possess spectral selectivity from e.g. 1.6; glazings without coatings and tinting from about 1.05 to 1.1. Glazings with S 1 are designated as non-selective.

Transparent Building Elements

114

1
External reflexion

Transmittance or reflectance (-)

visible
0.8

Internal reflexion
0.6

0.4

0.2
Transmission
0

0.5

1.5
Wavelength (m)

2.5

Fig. 6.8: A solar control glazing selectively transmits solar radiation [6.5]

6.3

Thermal Properties of Glazings

In order for the heat loss in a building to not be too large, windows must also exhibit, in addition to
good optical properties, good thermal properties, i.e. a thermal resistance as high as possible. The
heat flow through windows can be divided into two areas: (i) glazing and (ii) rim (spacer and
frame). The heat flux in the rim region is in general higher than in the glazing, which is why the area of the frames should be kept low.
Convection in gap
kalt
Heat conduction in glass pane
Convective heat transfer
to the exterior
Radiation exchange
with the exterior

warm
Heat conduction in gas filling
Convective heat transfer
to the interior
Radiation exchange
with the interior
Radiation exchange in the
gas space

Heat conduction in spacer

Fig. 6.9: Heat transport mechanisms in a glazing


The heat transfer in the gap between the glass panes takes place due to heat conduction, convection, and radiation (Fig. 6.9). The heat transfer coefficient for convection and conduction (c.f. section 1.2.3) can be expressed as follows:

Transparent Building Elements

115

Nu( Ra,Pr)
(W/m2K)
d
d denotes the distance between the panes, the conductivity of the gas filling and Nu the Nusselt
Number (Nu = dimensionless number that indicates how many times larger the heat transport due
to convection and heat conduction is compared with heat conduction only; see [6.7]). The heat
transfer coefficient for thermal radiation between the two glass surfaces amounts to (c.p. section
1.3.2):

gas =

1
T1 + T2

2 1 / 1 + 1 / 2 1
3

r = 4

(W/m2K)

T and denote the temperatures (unit: Kelvin) and emissivities, respectively, of both glass surfaces
in the gap. The Stefan-Boltzmann-constant is s = 5.6710-8 Wm-2K-4.
The total heat transfer coefficient in the gap is given as:

tot = gas + r

(W/m2K)

The thermal transmittance of a glazing Ug additionally contains the contribution of the internal and
external heat transfer resistances plus the thermal resistances of glass panes 1 and 2:
1
1 d1
1
d
1
=
+
+
+ 2 +
U g hi 1 tot 2 he

(m2K/W)

The thermal resistance of the glass panes d1/1 and d2/2, respectively, are small in comparison to
the other quantities. By replacing the air in the gap with a gas with better properties (noble gases
such as argon, krypton and xenon), the thermal insulation properties of the glazing can be improved (Tab. 6.1).

Air
Thermal conductivity (W/mK)
2

Argon
1.68410

-5

1.91010

-5

1.27410

2.49610

Thermal diffusivity a (m /s)

2.01010

Kinematic viscosity (m /s)

1.42910

Krypton

-2

-2

0.90010

-5

1.03210

-5

0.67410

Xenon

-2

0.53310

-2

-5

0.57510

-5

0.37710

-5
-5

Table 6.1: Thermal conductivity, thermal diffusivity and kinematic viscosity for air and different
gases at 10C [6.7, 6.8]
Figure 6.10 shows the influence of different filling gases on the heat transport by convection and
conduction in a gap as a function of the distance between panes. With small distances between
panes the heat loss due to conduction is dominant. With larger distances between panes, due to a
temperature difference in the glazing, the gas on the warm side rises up and sinks down on the
cold side (convection). Depending on the physical properties of the gas, the convection is initiated
at different distances between panes and a different heat flow results with a given distance between panes.
Figure 6.11 shows the influence of one or two low-emissivity coating(s) on the radiative heat transfer in the gap. Today low-emissivity coatings with < 0.05 are commercially available. If both glass
surfaces in the gap are coated in this way, the heat transfer coefficient for thermal radiation will
reach values of about 0.1 W/m2K (Fig. 6.11). In contrast, the heat transfer coefficient for convection
and conduction, even with a krypton- or xenon-filling, cannot be brought significantly below 1
W/m2K. Therefore, in modern glazings the heat transport through the gas filling is the dominant
mechanism for heat loss.

Transparent Building Elements

116

Air
Ar
Kr
Xe

Heat transfer coefficient

gas

(W/m K)

3
2.5
2
1.5
1
0.5

DT = 15 K

0
0

0.01

0.02

0.03

Distance between panes (m)

Fig. 6.10: Heat transfer coefficient for convection and conduction gas in a gap as a function of the
distance between panes for different gas fillings ([6.9], calculated according to [6.7])

Hat transfr cofficint (W/m K)

0.5
On pan coatd
(uncoatd = 0.84)

0.4

0.3
0.2
Two pans coatd

0.1
0
0

0.02

0.04

0.06

0.08

0.1

Emissivity , (-)
1

Fig. 6.11: Heat transfer coefficient for thermal radiation in a gap between two glass panes as a
function of the emissivities of the surfaces at 10C [6.9].
Table 6.2 gives numerical values for Ug and g as well as the transmittance for visible light V of
glazings. Comparing to an older air-filled double glazing with a Ug = 3 W/m2K, the heat that flows
with a given temperature difference through a modern triple glazing with a Ug = 0.5 W/m2K is only a
sixth!
The spacers at the edges hold the glazing together mechanically and prevent the leakage of the
filling gas (Fig. 6.12). The heat conduction through the spacer however increases the heat flux at
this location. In winter, due to lower surface temperatures, if condensation occurs on the internal
surface of the glazing, this is the most likely location.

Transparent Building Elements

117

In calculating the heat flow through a glazing, the thermal bridge effect of the spacer is evaluated
with a linear thermal transmittance (Table 6.3).

Ug (W/m K)

v (-)

g (-)

v (-)

g (-)

Ug (W/m K)

Table 6.2: Values for Ug, g and v of glazings [6.3](* with 10 % air)
glass
spacer

primary seal

desiccant

secondary seal

Fig. 6.12: Cross-section of the edge region of an insulating glazing [6.10].

Glazing

Spacer (W/mK)

Ug (W/m2K)

Table 6.3: Values for the linear thermal transmittance of different spacers [6.3]
Because glazings, spacers and window frames possess different insulation properties, these three
regions must be considered separately in the calculation of the thermal transmittance of windows.
Figure 6.13 shows a cross-section of a window in an exterior wall and the parameters to calculate

Transparent Building Elements

118

the global thermal transmittance of the window Uw. The thermal transmittance of the window can
now be calculated as follows:
Uw =

U g Ag + U f Af + L
Aw

(W/m2K)

Ug

Thermal transmittance of the glazing

W/m2K

Uf

Thermal transmittance of the frame

W/m2K

Linear thermal transmittance of the glazing edge

W/mK

Ag

Projected area of the glass

m2

Af

Projected area of the frame

m2

Length of the glazing edge (spacer)

Aw

Projected area of the window (Aw = Ag + Af)

m2

Wall wall interface

- glazing edge

Uf

Ug

Fig. 6.13: Cross-section through a window in a wall opening

Frame
Uf
2

(W/m K)

Table 6.4: Uf values for different window frames [6.3]


In addition, the interface between the window and the wall is accounted for with a linear thermal
transmittance Wall.

Transparent Building Elements

6.4

119

Energy Fluxes Through Windows

The greater the thermal transmittance U and the greater the temperature difference between the
interior and exterior is, the greater will be the transmission heat loss through the glazing. The solar
energy flux produces a heat gain through the glazing and will be greater the greater the total solar
energy transmittance g is and the greater the solar irradiance I is.
If U > g I :

Net energy loss

If U = g I :

Equilibrium (loss = gain)

If U < g I :

Net energy gain

Which situations arise now with different glazings in varying climates and according to the facade
orientation? The energetic quality of a glazing can be expressed by the ratio g/U (c.p. Fig. 6.3) and
the exterior climate by the quotient /I. The interior temperature is assumed to be 20C. Figure
6.14 shows that in Zurich in December (old) air-filled and uncoated double glazings at all faade
orientations produce mean monthly heat losses. Because the solar irradiance on the north facade
is the least, the net losses here are the greatest. In contrast a modern triple glazing exhibits, especially on the south but also on the west and east facade, a positive monthly balance. In Figure 6.14
denotes the gain-to-loss ratio:

1.4
1.2

West

Est

=1

=8
=2
=4

Loss

Gin

Triple glzing
2
U = 0.4 W/m K, g = 0.47

= 1/2

0.8

ouble Glzing
2
U = 1.0 W/m K, g = 0.60

g/U (m K/W)

North

g I
U
South

0.6
0.4

= 1/4

0.2

= 1/8

ouble Glzing
U = 2.9 W/m2K, g = 0.77

= 1/16

0
0

0.5

1.5

Single glzing
2
U = 5.9 W/m K, g = 0.87

/I (m K/W)
2

Fig. 6.14: Net gain and loss, respectively, through different glazings with different faade orientations [6.15] (climate data for Zurich-SMA, monthly mean values for December [6.11])

6.5

Solar Shading Devices

Solar shading is understood as all methods that inhibit an over-heating of the interior from solar irradiation. Solar shading devices can be either fixed with respect to time, i.e. unchangeable, or var-

Transparent Building Elements

120

iable. The solar shading can be placed at different locations: at the exterior, integrated in the window or at the interior (Table 6.5).

Variability with time


Variable

Location
Exterior

Fixed

- textiles
- venetian blinds
- roller blinds
- textiles
- venetian blinds
- electrochromic devices
- curtains
- venetian blinds

Integrated in window

Interior

- overhangs
- solar control glazing

Table 6.5: Classification of solar shading devices


External shading devices have the great advantage in that they can give off the absorbed solar energy directly to the exterior surroundings (Fig. 6.15). With internal shading devices the absorption
of the solar irradiance occurs within the thermal insulation. This is with regards to over-heating of
the interior very disadvantageous. With external shading devices, therefore, significantly lower
total solar energy transmittances can be achieved compared with internal ones.

Internal shading

Temperature

Temperature

External shading

Fig. 6.15: Energy flows and temperature profiles with external and internal shading devices
Fixed shading devices such as overhangs use the seasonal differences of the suns position: the
suns altitude is small in winter and large in summer. These solar shading devices are however especially suitable to screening direct solar irradiance on south facades. Their effect is limited with
east and west facades and with respect to diffuse radiation. Fixed shading devices furthermore often block a significant part of the daylight (zenith light), also during times when the solar irradiance
is low, which is of course very disadvantageous.
In addition to the light transmittance v, the total solar energy transmittance g is the most important
property of a shading device. The total solar energy transmittance g applies to a layer sequence,
e.g. for a triple glazing with external venetian blinds. Not only the optical properties of all the layers
but also, for example, the thermal resistance of the glazing (U-value) influences the resulting gvalue. The total solar energy transmittance indicates which portion of the incident solar energy with
closed solar shading will accumulate as heat in the interior space. Total solar energy transmittances g < 0.15 are easy to achieve with external or integrated, well (possibly mechanical) ventilated,
solar shading devices and should be strived for.

Transparent Building Elements

121

Due to the multitudes of possible combinations it is reasonable in critical cases (e.g. with highly
glazed office buildings), to determine the g-value of a transparent building element either experimentally (solar calorimeter) or numerically. Because the optical properties of the solar shading devices and glazings are in some cases strongly wavelength dependent, a calculation of the total solar energy transmittance (and the light transmittance) must be carried out, in general, wavelength
dependent. Thus a solar control glazing can, for example, transmit well visible radiation, but be virtually opaque for infrared radiation (c.p. e.g. Fig. 6.8). Slats of venetian blinds can e.g. have a similar solar reflectance but a very different visible reflectance (e.g. beige w.r.t. dark red in Table 6.6).
Specialized software for planning purposes is available for the spectral calculation of v and g that
also contain data bases with numerous commercially available glazings and solar shading devices
(e.g. [6.12]). With glass double-skin facades the occurring airflow patterns are important to the resulting total solar energy transmittance [6.13, 6.14].

Color

Solar

Visible

UV

Color

Solar

Visible

UV

yellow VSR-720

0.552

0.493

0.068

light green 3040-G

0.274

0.320

0.071

white VSR-010

0.742

0.837

0.084

bronze VSR-780

0.252

0.249

0.155

grey VSR-130

0.392

0.461

0.079

dark green VSR-220

0.185

0.097

0.068

aluminum VSR-140

0.489

0.490

0.549

dark blue VSR-440

0.271

0.130

0.069

light beige VSR-240

0.585

0.575

0.087

dark red VSR-330

0.356

0.092

0.062

beige VSR-110

0.327

0.342

0.081

black 8505

0.064

0.065

0.063

Table 6.6: Weighted reflectances in three wavelength intervals for different colors of slats of venetian blinds [6.5]
In the selection of a suitable system from the many possible combinations of glazings and solar
shading devices in different types and sequences (external/integrated/internal) and varieties of
ventilation (natural/mechanical), one should bear the following in mind:
- An external, variable solar shading device is basically the best solution, if the costs, demands
and architectural expression allow for it.
- The further inside a solar shading device is located, the more energy will be transmitted into the
interior. From the shading device, absorbed energy can more easily flow into the interior, the
lower the thermal resistance between the shading device and the interior is compared with the
thermal resistance between the shading device and the exterior ( U-value glazing).
- Solar shading devices fully or partially made of opaque materials can also serve as glare
shields. However conflicting demands can arise.
- Venetian blinds (external, integrated or internal) also allow, to a certain extent, a redistribution of
the daylight (light guiding due to reflection on the slats).
- The selection of shading device and glazing is also significant with regards to the resulting internal surface temperatures and thermal comfort (chapter 3). Highly absorbing layers, especially
if arranged on the internal side, have an adverse effect in summer.
- With wind exposed situations as with high-rise buildings, buildings in the mountains, etc. the
wind pressure on the shading device must practically always be reduced with external glass
panes: integrated solar shading naturally ventilated to the exterior or with a mechanical ventilation system or solar control glazing.
- Electrochromic layers as shading devices, i.e. smart glass (switchable), are still in the developmental stages but could become more meaningful in the long term (costs, switching hub, service life, appearance).

Transparent Building Elements

122

Problems
Problem 1: Energy fluxes with glazings in winter
Given are two glazings:
(i)

Double glazing: Ug = 1.2 W/m2K and g = 0.61

(ii) Triple glazing: Ug = 0.5 W/m2K and g = 0.45


The external and internal temperatures are e = -5C and i = 20C, respectively. Through these
glazings the interior gains energy due to solar radiation, but also loses energy due to heat transmission.
a)
b)
c)

Formulate the energy fluxes assuming an equilibrium of gains and losses.


How great is the critical solar irradiance under these conditions, above which a net energy
gain for the interior results?
With what weather conditions can this irradiance be reached?

Problem 2: Temperature of a solar control glass


A glass pane (e = 0.52) is placed as an apron in front of a facade. Calculate the temperature of
the glass with the following boundary conditions:

e = 35C, he = 12 W/m2K, I = 600 W/m2, i = e and hi = 10 W/m2K.

Problem 3: Heat transport in a double glazing


Given is a double glazing with a krypton gas filling. The thicknesses of the glass panes are 6 mm
each and they are spaced 10 mm apart. Both inner glass surfaces are coated (1 = 0.04 and 2 =
0.08) and gas = 1 W/m2K can be taken for the heat transfer coefficient in the gap. The thermal
conductivity of the glass is = 1 W/mK and the heat transfer coefficients at the glazing surfaces
can be taken as he = 25 W/m2K and hi = 7.7 W/m2K, respectively.
a.) Draw the thermal flow diagram for the glazing including all thermal resistances.
b.) Calculate the thermal transmittance coefficient Ug, assuming that the average surface temperature in the gap is 5C.
c.) Calculate the percentage of the thermal resistances heat transfer at surfaces, glass panes
and gap assuming that the total resistance is 100%.

Problem 4: Thermal comfort next to a glazing in summer


A double glazing possesses the following properties:
- thermal transmittance Ug = 1.3 W/m2K
- solar absorptance in the external pane 1 = 0.07
- solar absorptance in the internal pane 2 = 0.28
The boundary conditions are given by:
- Solar irradiance I = 600 W/m2
- external air temperature e = 33C und internal air temperature i = 26C
- heat transfer coefficients he = 20 W/m2K and hi = 8 W/m2K
Calculate the temperature of the internal pane and interpret your result.

Transparent Building Elements

123

Problem 5: Solar protection at an exposed location


A restaurant with a mountain vista is planned in the alps. At this location, the external temperature
can go down to -20C and the solar irradiance reaches values up to 900 W/m2. Due to wind exposure an external shading system is not possible. Please comment on which and how parameters of
the glazing must be chosen in order that
a)

with low external temperature and little solar irradiance, the internal surface temperature lies
as close as possible to the room temperature?

b)

the glazing warms up only a little if the solar irradiance is high?

c)

behind the glazing with high solar irradiance (direct sunshine!) one could still eat comfortably?

Problem 6: Secondary internal heat transfer factor of a single glazing


Calculate the secondary internal heat transfer factor qi of a single glazing as a function of the internal and external heat transfer coefficients (hi and he respectively) as well as the solar absorptance
of the glass pane e:
a.) As a generally valid formula (hint: apply the conservation of energy, the definition of qi as well
as the assumption internal temperature = external temperature)
b.) with he = 25 W/m2K, hi = 7.7 W/m2K and e = 0.1

Literature
[6.1]

Platzer W., Wittwer V., Transparent Insulation Materials, Chapter 3, in: Materials Science
for Solar Energy Conversion Systems, Ed. Granqvist C.G., Pergamon Press, Oxford, 1991

[6.2]

Goetzberger A., Wittwer V., Sonnenenergie: Thermische Nutzung, Teubner, Stuttgart, 1989

[6.3]

Zrcher C., Frank T., Bauphysik: Bau und Energie, vdf, Zrich, 2004

[6.4]

Duffie J.A., Beckman W.A., Solar Engineering of Thermal Processes, John Wiley & Sons,
New York, 1991

[6.5]

Manz H., Frank T., Thermal simulation of buildings with double-skin faades, Energy and
Buildings, Vol. 37, 2005, 1114-1121

[6.6]

EN 410, Glas im Bauwesen Bestimmung der lichttechnischen und strahlungsphysikalischen Kenngrssen von Verglasungen, Europisches Komitee fr Normung, Brssel, 1998

[6.7]

EN 673, Glas im Bauwesen Bestimmung des Wrmedurchgangskoeffizienten (U-Wert)


Berechnungsverfahren, Europisches Komitee fr Normung, Brssel, 1997

[6.8]

ISO/DIS 15099 (Draft), Thermal Performance of Windows, Doors and Shading Devices
Detailed Calculations, ISO Central Secretariat, Geneva, Switzerland, 2003

[6.9]

Manz H., On minimizing heat transport in architectural glazing, Renewable Energy, Vol. 33,
2008, 119128

[6.10] Platzer W. J., Fenster und Verglasungen, in: Thermische Solarenergienutzung an Gebuden (Hrsg: Marko A., Braun P.), Springer-Verlag, Berlin, 1997
[6.11] SIA 381/2, Klimadaten zu Empfehlung 380/1 Energie im Hochbau, Schweizerischer Ingenieur- und Architekten-Verein, Zrich, 1991
[6.12] GLAD, Glasdatenbank und Rechenprogramm, Eidgenssische Materialprfungs- und Forschungsanstalt, Dbendorf, 2008
[6.13] Manz H., Schaelin A., Simmler H., Airflow patterns and thermal behavior of mechanically
ventilated glass double faades, Building and Environment, Vol. 39, 2004, 1023-1033

124

Transparent Building Elements

[6.14] Manz H., Total solar energy transmittance of glass double faades with free convection,
Energy and Buildings, Vol. 36, 2004, 127-136
[6.15] Manz H., Menti U.-P., Energy performance of glazings in European climates, Renewable
Energy 37 (2012) 226232

Chapter 7
Air Exchange

Air Exchange

126

Air Exchange

Due to wind and/or temperature differences, pressure differences can occur between the interior
and exterior that produce airflows through the air leakages in the building envelope. If the windows
and doors are closed the air exchange will particularly occur through joints in the building envelope. This unintentional air exchange is called infiltration (= airflow to interior) and accordingly exfiltration (= airflow to exterior). The air exchange through open windows will be referred to as natural
ventilation and that with an air-handling system as mechanical or controlled ventilation.
The air exchange in an interior room must satisfy requirements in the following areas:
Indoor air quality: low concentration of pollutants (dust, allergens, smoke, etc.) and odors
Thermal comfort: air velocity and turbulence not too high, air temperature and humidity within
the comfort range
Building construction and humidity: prevention of mold growth (removal of humidity that is produced in the interior) and building damage (danger of moisture penetration into building components at leakages)
Heat loss: minimize the energy consumption for heating
With regards to air quality and the removal of moisture, a high air change rate is advantageous.
For thermal comfort and to keep the heat losses small, a low air change rate is demanded, where
the requirements relating to heat losses are significantly stricter. The requirements listed above are
partly conflicting as regards the air change rate in a room. One is confronted with a dilemma!
On the grounds of indoor air quality and to remove moisture, about a half of the air volume in an
interior room must typically be replaced per hour in residential buildings. The heat loss due to the
air exchange is secondary with poorly insulated buildings. If the thermal insulation is improved then
the heat transmission losses decrease, i.e., that relatively the ventilation losses gain significance. If
the building is well insulated, then the ratio of the ventilation heat losses to the total heat loss increases to more than a half (Fig. 7.1): The ventilation heat losses become larger than the transmission heat losses!

Ratio of ventilation heat


losses to total heat losses (-)

Insulation level:
High
Average

Low

Air change rate (1/h)

Fig. 7.1: Ratio of ventilation heat losses to total heat losses (total = transmission + ventilation) in a
building [7.1]
For buildings in a climate with cold winters as in Switzerland, the solution to the dilemma concerning the air exchange described above is:
(i)

The building envelope shall be as air-tight as possible [7.2].

Air Exchange

(ii)

127

The necessary air exchange (in winter) should be provided for by mechanical ventilation with
heat recovery (Section 7.6).

In addition to the already mentioned features the ventilation can additionally be used to cool a
building at night to keep it comfortable in summer (Section 7.7).

7.1

Wind Pressure on the Building Surface

When wind blows against a building, the pressure distribution on its surface changes. This topic
will be considered in the following and the most important phenomena will be presented.
Bernoullis equation (Daniel Bernoulli, 1700-1782) describes a frictionless, incompressible, steadystate flow. Such an idealized flow in the vicinity of an obstacle is shown in Figure 7.2.

Reference Point

Streamline

v1, p1

v0, p0

h0

h1

Fig. 7.2: Idealized flow around a cylinder


According to Bernoulli, for two points on the same streamline the following applies:
1
1
v 02 + g h0 + p0 = v12 + g h1 + p1
2
2

p0
v0
p1
v1

static pressure at reference point (free, undisturbed flow)


velocity at reference point (reference velocity)
static pressure at a given point
velocity at a given point
density of air
gravitational acceleration (g = 9.81 m/s2)
height

Pa
m/s
Pa
m/s
kg/m3
m/s2
m

g
h
1
v2
2
g h

dynamic pressure

Pa

gravitational pressure

Pa

static pressure

Pa

For small height differences as e.g. with buildings the gravitational pressure can approximately be left out. Thus through rearrangement of Bernoullis equation the pressure difference between
the reference point 0 and a given point 1 is obtained:

p = p1 p0 =
v1 < v0 :

1
v 02 v12
2

p > 0

positive pressure (p1 > p0), with v1 = 0: p =

1
v 02 (stagnation pressure)
2

Air Exchange

128

v1 > v0 :

p < 0

negative pressure (p1 < p0)

Velocity field and pressure distribution are coupled together. According to the flow pattern and the
occurring velocities, in certain areas of the building envelope a higher pressure and in other areas
a lower pressure will result.
In reality airflow around buildings also give rise to other phenomenon such as flow separation and
turbulence that cannot be described by Bernoullis equation (Figures 7.3 and 7.4). On the upstream flow side (= windward) the Bernoulli equation approximately gives good values, however
not for the backside (= leeward side).
flow separation

wake

Fig. 7.3: Flow separation with real flow und leeward side wake zone

Fig. 7.4: Visualization of streamlines and eddy formations with smoke from air flowing diagonally
against a square building in a wind tunnel test [7.3]
The flow around buildings is particularly dependent on:
wind velocity and direction
terrain properties in the vicinity of the building (wind profile see chapter 2)
geometry of the building (basic shape, roof inclination, overhangs and protrusions)
neighboring buildings
Figure 7.5 schematically shows the airflow around a rectangular building. In front of the building an
eddy develops. Sharp corners and edges give rise to flow separation and wake flows behind the
building.

Air Exchange

129

Fig. 7.5: Airflow around a rectangular building (left: side view, right: perspective depiction)[7.3]
Figure 7.6 schematically shows the pressure distribution on the surface of a building due to a perpendicular oncoming airflow. On the windward side of the facade a positive pressure occurs and a
negative pressure on the roof as well as the side and leeward facade. The wind profile (cp. chapter
2) is dependent on the terrain properties (surface roughness) in the vicinity of the building. Different
wind profiles give rise to different pressure distributions on the building surface.

Fig. 7.6: Schematic pressure distribution on the surface of a building due to a perpendicular oncoming airflow. (above: side view, below: top view) [7.3]
The occurring pressure on the building envelope is characterized by the pressure coefficient Cp.
These coefficients refer to the dynamic pressure of the undisturbed airflow (stagnation pressure).
The static pressure on the building envelope is designated by pF.

pF p0 = C p

1
v 02
2

Cp-values for typical building geometries and different airflow directions are available in the literature. In utilizing these mostly experimentally determined values it is however necessary to use caution. Among other things, because e.g. from neighboring buildings the incoming flow conditions on
real buildings can deviate from the situations in experiments and thus develop different pressure
distributions on the building surface.

Air Exchange

130

Gap effect

Nozzle effect

Redirection effect

Fig. 7.7: Airflows between buildings


If two or more buildings are close to each other new flow patterns can develop compared with the
case of a building without any neighboring buildings. Figure 7.7 shows some flow effects between
buildings. With complicated situations wind tunnel tests can depict the expectable flow pattern (Fig.
7.8). With larger building projects such tests are sometimes conducted. The laws of similarity must
be considered in wind tunnel experiments. In recent years research on airflow around and in buildings has increased with the help of simulations ( computational fluid dynamics = numerical fluid
mechanics).

Fig. 7.8: In a wind tunnel test the streamlines that form with a group of buildings can be made
visible with smoke.
Figure 7.9 shows the calculated frequency distribution of wind-induced pressure on a building facade. The wind velocity v0 was taken from a weather data file [7.5] for the not very wind exposed
location of Zurich-SMA; with a density = 1.2 kg/m3 and pressure coefficient Cp = 0.1 / 0.5 / 1, respectively. The figure shows that in Zurich the wind-induced pressure is smaller than 2 Pa about
50% of the time. However, wind-induced pressures can often rise up to tens of Pascals.

131

50

pF p0 = C p

40

1
v 02
2

C =1
p

30

C = 0.5
p

20

Cp = 0.1

10

99.9

99

5
10
20
30
50
70
80
90
95

0
.1

Wind-induce pressure on the facade (Pa)

Air Exchange

Fraction of time (%)

Fig. 7.9: Calculated frequency distribution of wind-induced pressure on a building facade as a


function of the pressure coefficient Cp at location Zurich-SMA (height 10 m above ground)[7.4]

7.2

Thermally Induced Pressure Differences

The air pressure p decreases exponentially with increasing height above ground in the still, isothermal atmosphere:
p(h ) = p0 e

p0

0
g
h

0 g h
p0

pressure at ground level


density at ground level
gravitational acceleration
height above ground

For small height differences, such as e.g. height differences in buildings, the equation can be linearized:
p(h ) p0 0 g h

The pressure difference p between a reference height with an average pressure pm (z = 0) and
height z thus amounts to:

p( z ) = pm p( z ) = 0 g z
The temperature of the interior and exterior air and consequently also the density are often different, so that the pressure increase from below to above proceeds differently.

pi ( z ) = i g z

respectively

pe ( z ) = e g z

Applying the ideal gas law (p = RT), which describes the interrelationship between pressure,
density and temperature, one gets for the pressure difference pie between the interior and exterior:

pie ( z ) = pe ( z ) pi ( z ) = pm

1 1
1
g z
R
Te T i

Air Exchange

132

gas constant for air

R = 287.1 J/kgK

acceleration of gravity

g = 9.81 m/s2

pm

atmospheric air pressure

pm 95000 Pa

Based on the reference height (z = 0), where the neutral zone with pi = pe = pm is located, in buildings with Ti > Te a linear upwardly increasing positive pressure occurs with increasing coordinate z
(Fig. 7.10). In halls, open stairwells (large heights) and chimneys (large temperature differences)
considerable pressure differences can occur.
Stack effect

cold

Pressure profiles

warm

Fig. 7.10: Pressure profiles interior (warm) and exterior (cold)[7.6]


The spatial distribution of leakages in the building envelope influences these pressure differences
between the interior and exterior. With a vertically uniform distribution of air leakages the neutral
zone is located in the middle of the facade height. If one leakage is very much larger than all others, the neutral zone will shift towards this location.
Figure 7.11 shows thermally induced pressure profiles in multi-story buildings. In case A a large
permeability exists between the individual stories and the pressure develops uninterrupted over the
entire facade height. With full air-tightness between stories (case B), only the height of a story can
be effective. With actual buildings (case C) mostly a mixture of both limit cases A and B is encountered. It consists of neither full permeability nor complete tightness between stories. With regard to
the total airflow resistance in a building interior, vertical shafts, e.g. staircases, are especially important. The flow resistance is here much smaller than with the air leakages between stories.
Figure 7.12 shows a frequency distribution of the thermally induced pressure difference between
the interior and exterior. The external temperature Te is taken from weather data for Zurich-SMA
and the internal temperature Ti at 293 K or 20C. The thermally induced pressure differences occurring in Zurich are mostly smaller than about 10 Pa.

Air Exchange

A. Permeable floor levels

133

B. Air-tight floor levels

C. Permeable floor levels with staircase

15
z = 10 m
10
z=5m
5
z=1m

99.9

99

5
10
20
30
50
70
80
90
95

-5

.1

Thermally induced pressure differences


between interior and exterior (Pa)

Fig. 7.11: Pressure profiles for multi-story buildings with varying leakage characteristics [7.6]

Fraction of time (%)

Fig. 7.12: Cumulative frequency distribution of thermally induced pressure differences between the
interior and exterior as a function of the distance from the neutral zone at the location Zurich-SMA
[7.4]

7.3

Airflow Through Leakages

Due to wind (Section 7.1) or temperature differences (Section 7.2) pressure differences occur over
the building envelope. In comparison to the atmospheric air pressure, approximately 95'000 Pa in
Zurich, these pressure differences are very small (Fig. 7.13).

Air Exchange

134

Wind speed v (m/s)


0

Pressure difference p (Pa)

20
z = 10 m

Wind-induced pressure difference:


1
p = C p v 02
2

15
Cp = 1
10

z=5m

Thermally induced pressure difference (i = 20C):


1 1
1
pie = pm g z
R
Te Ti

5
z=1m
0
-20

-10

10

20

External air temperatur (C)


e

Fig. 7.13: Thermally and wind-induced pressure differences (i = 20C)


Air leakages in the building envelope can occur with window or door joints, connections (e.g. wallfloor or wall-roof), ductwork (installations), or built-in components (window frames, roller shutter
housings). The airflow through such joints or gaps can be described as follows:
m
V = D (Dp )

air flow rate

m3/h

coefficient, characterizing permeability

m3/hPam

pressure difference over the leakage

Pa

exponent, dependent on the type of flow


m = 1: fully laminar; m = 0.5: fully turbulent

The flow-exponent is often taken as m = 2/3, because actual flows through leakages are generally
neither fully laminar nor fully turbulent. The permeability coefficient D for a joint can be described
as follows:
D = al

air leakage coefficient

m3/(hmPam)

joint length

The air-tightness of windows and doors can be determined in a laboratory experiment as a function
of the different pressures [7.7]. Todays window construction exhibit air leakage coefficients from
about 0.01 to 0.04 m3/(hmPa2/3), older windows and doors with poor sealing or no gaskets possess air leakage coefficients from about 0.2 to 0.6 m3/(hmPa2/3). Figure 7.14 shows the influence
of the air leakage coefficient on the airflow rate per meter of joint length that arises with a given
pressure difference.

Air Exchange

135

V
2/3
= a (p )
l

10

2/3

2/3

a = 0.5 m /(h m Pa )

joint length (m /h m)

Air flow rate per meter of

12

6
4
a = 0.1 m /(h m Pa )

20

40

60

80

2/3

a = 0.02 m /(h m Pa )
100

Pressure difference p (Pa)

Fig. 7.14: Air flow rate per meter of joint length as a function of the pressure difference and air
leakage coefficient
The wind-induced air exchange in a building with two openings can be modeled as shown in Figure
7.15. This arrangement can be considered with an analogy to electrical circuits as a series connection of two resistors. According to the above equation it gives:
m
V1 = D1 (p1 pi )
m
V2 = D2 (pi p2 )

The continuity equation (inflow = outflow) stipulates that


V1 = V2

If the pressures on the facade, p1 and p2, as well as the parameters to describe the leakage flows,
D1, D2 and m, are known then from the three equations the three unknowns, the interior pressure pi
and the airflow volumes, V1 and V2 respectively, can be calculated. Contrary to Ohms law for electrical circuits the equations are however nonlinear (m = 2/3).

Fig. 7.15: Wind-induced air exchange in a building with two openings


Problems with multiple openings and zones (= areas with constant pressure) can be solved analogously. An example for a multi-zone scheme is shown in Figure 7.16. In addition to the airflow resistances in the building envelope resistances also occur in the building interior (doors). In each
room a different interior pressure arises.

Air Exchange

136

Fig. 7.16: Floor plan with corresponding network [7.6]

7.4

Indoor Air Quality

The indoor air should not contain any substances hazardous to health and should feel fresh and
comfortable. By the term indoor air quality one means the non-thermal aspects of the interior air
that is relevant to both the well-being and health of people. Pollutants in the indoor air can stem
from different sources:
Polluted outside air
Sources of air pollution in the interior
Radon in the ground
Especially in urban areas (traffic, heating, industry), the outside air has certain pollutants. Due to
allergic reactions pollen can also, for example, be considered a pollutant. With mechanically ventilated buildings there is the possibility to reduce the concentration of certain pollutants in the supply
air with suitable filters. It makes most sense of course to avoid having pollution sources and not to
have to provide excessive ventilation. For the interior rooms this means that construction materials
and above all interior finishes (timber products, joint sealants, floor coverings, etc.) should be used
that emit only very limited or no pollutants (solvents, fungicides, formaldehyde, etc.). A careful
choice of materials based on clear product declarations can possibly avoid a lot of frustration and
health ailments.
Certain pollutants cannot be avoided with the use of rooms. These substances must be removed
by exchanging the air. In order to remove the exhaled carbon dioxide and body odors, a supply of
fresh air of about 15 to 30 m3/h per person is recommended. As a relatively simple measurable
quantity that characterizes the freshness of the room air, the concentration of carbon dioxide is
generally used. As an upper limit of the hygienic range 1500 ppm (= 0.15 Volume-%) is recommended. The CO2-concentration of the external air amounts to just under 400 ppm.
Radon is a radioactive gas that can be emitted from certain rocks. The main source of radon is the
ground below a building. The level of radon emissions depends largely on the geologic formation,
which is why the radon emission throughout Switzerland is very different. High levels are particularly found in the regions of Graubnden, Ticino and Jura (Fig. 7.17).

Air Exchange

137

BAG 2007

Fig. 7.17: Radon map of Switzerland based on interior room measurements [7.8]
According to the Swiss Federal Office of Health (2008) 200 to 300 people die in Switzerland each
year from radon caused lung cancer. Hence radon is the most dangerous carcinogen in residential
areas and is after smoking the greatest cause for lung cancer. The risk of lung cancer is greater
the higher the radon concentration in the inhaled air is and the longer one breathes in this air.
Whether radon can penetrate into a house depends primarily on how tight the house foundation
against the ground is. Especially disadvantageous therefore are buildings without a foundation slab
made of concrete but with e.g. a pebble stone floor (high permeability between ground and indoor
spaces), which otherwise possess a relatively tight building envelope. The influx of radon is thereby possible, but the exfiltration is hampered so that an accumulation is possible.
The most important measures to hinder the harmful radon accumulation in the interior rooms are
therefore: (i) to minimize the radon entry from the ground by sealed basements, i.e. foundation slab
made of concrete and possibly a film, (ii) airtight floors separating the cellar and living space stories, as well as (iii) avoiding negative pressure in the building interior.

7.5

Airtightness of the Building Envelope

Thermally or wind-induced pressure differences over openings such as joints, open windows and
doors produce an exchange of interior air with exterior air. This air exchange is also influenced by
the building design (properties of the building envelope, arrangement of the interior rooms) and the
building operation (opening/closing of windows and doors as well as possible use of air extract
units). With a mechanical ventilation system pressure is produced on the room openings by fans
that cause an air exchange.
Generally effective is that the building envelope which encloses the heated volume be as tight as
possible. The required quantity of external air is ensured through the manual opening of the windows, other ventilation openings or through air-handling units [7.2].
The interior air is in general not still but is propelled by different causes:
- Upward and downward drafts result from temperature differences (such as cold window surfaces, warm radiators, people and appliances, e.g. computers)
- Movement (people, fans, possible vehicles)
- Momentum of supply air (mechanical ventilation or window ventilation)

Air Exchange

138

However, air flow patterns in the room are not of further interest here but instead the global degree
for the exchange of interior air by exterior air. The so-called air change rate n indicates how often
the room air is exchanged per hour:
n=

V
V

(1/h)

air flow rate interior-exterior

m3/h

volume of room

m3

The airtightness of a building envelope can be experimentally determined by means of the pressure difference method (blower door test) [7.9]. For the measurement in the building envelope installed fans, mostly in a door, produce a pressure difference between the interior and exterior (Fig.
7.18). To eliminate weather influences as much as possible, this artificially produced pressure difference must be significantly larger during the measurement than the naturally occurring pressure
difference on the building envelope. The leakage airflow rate is measured as a function of the
pressure difference over the building envelope. This relationship can be described by the flow
equation for joints or gaps (section 7.3). The air change rate with a pressure difference of 50 Pa is
denoted by n50.

m
V = D (p )

p
p
V
Negative
internal pressure

Positive
internal pressure

Fig. 7.18: Pressure difference method to determine the air-tightness of a building envelope: principle (left) and a blower installed in a door (right)
If a negative pressure is produced in the building interior with the blower installed in the building
envelope at low external temperatures, the cooling of the inner surface of the building envelope
can be visualized with an infrared image and thereby indicate where the leakages can be found in
the building envelope.
With a carefully planned and executed building envelope an air-tightness value of n50 < 0.6 1/h (=
requirement for passive house standard [7.10]) can be reached. However, it is somewhat easier to
achieve lower values of n50 the more compact and large the building is, because the volume increases by the third power, while the envelope area increases only by the second power.
In the SIA standard 180 [7.2] the specific value used for the air-tightness of the building envelope
refers to the envelope area Ae and a pressure difference of 4 Pa:
v a,4 =

V4
Ae

m3

h m2

Air Exchange

139

The ratio between the two specific values va,4 and n50 depends on the building geometry:
m

v a,4
V 4
V
=

0 .2
n50
Ae 50
Ae

m denotes the flow exponent (section 7.3). Table 7.1 shows the maximum allowable value for the
air permeability of the building envelope according to SIA 180. For mechanically ventilated buildings the target values are to be met.

va,4 max (m /hm )


Limit value Target value
3

New construction

0.75

0.5

Renovation

1.50

1.0

Tab. 7.1: Limit- and target values for the air permeability of building envelopes [7.2].
To reach as much as possible the target of a continuously air-tight building envelope, the air tight
layer of the construction has to be carefully designed and appropriate materials have to be chosen
(i.e. concrete, plastered masonry, airtight membrane). With present day carefully constructed and
executed envelopes, general connection interfaces of every kind are where leakages dominate
(e.g. window/wall, wall/roof), especially also at ducts for electrical conduits (plugs) and pipes from
sanitation or other technical installations. Careful workmanship and intermediary inspections on the
construction site are recommended. The high air-tightness must not only be provided at the close
of construction but must also be preserved during the service life of the building.

7.6

Mechanical Ventilation with Heat Recovery

The natural ventilation of buildings is often not considered optimal, because due to wind and temperature conditions as well as not clearly defined leakages, unwanted air exchanges arise. In contrast a mechanical ventilation system can regulate the air exchange based on need. Many living,
work, and school spaces are however only used for a small amount of the time. Contaminants
need to be removed by ventilation only when someone occupies a room. If, e.g., a room is in use
eight hours a day on weekdays, then the average time use is only 24%!
It makes sense to equip an air-handling unit with the possibility for heat recovery whereby in winter
a significant portion of the ventilation heat losses (up to about 80 %) can be saved. Mechanical
ventilation with heat recovery is thus an essential measure in a climate with cold winters in providing buildings with a very low energy demand (e.g. Minergie or passive house). A prerequisite for
the application of these systems is a tight building envelope.
Generally there are many possible systems to ventilate residential buildings (Fig. 7.19). However,
for low energy buildings in climates with cold winters only ventilation systems with heat recovery
are suitable.

Air Exchange

140

Residential ventilation

Mechanical with heat


recovery

Natural

Window

Shaft

Heat exchanger:
recuperator
regenerator

Mechanical without
heat recovery

decentral

central

Supply and
extract air

decentral

central

Extract air

Supply and
extract air

Supply and
extract air

Extract air

Heat exchanger
and heat
pump

Heat exchanger:
recuperator
regenerator

Heat exchanger
and heat
pump

Extract air

Heat pump

Fig. 7.19: Systems for residential ventilation [7.11]

Ventilation system with heat recovery


Ground
heat exchanger

Fig. 7.20: Residential ventilation system with a ground heat exchanger and heat recovery [7.10]
A typical residential ventilation system for low energy buildings is displayed in Figure 7.20. In addition to the controlled supply and extract air flows and the heat recovery, a ground heat exchanger
is often used. A ground heat exchanger consists of nearly horizontal pipes that are buried in the
ground about 1.5 m to 3 m deep around the building excavation or also under free surfaces (e.g. a
garden or a parking lot). In winter the cold outside air flows through these pipes in the warm ground
(cp. section 2.3), whereby a preheating of the fresh air is obtained. In summer the ground heat exchanger can also be used for cooling the air as needed, because the ground temperature is then in
general considerably below the exterior air temperature.

Air Exchange

141

In a mechanically ventilated housing unit it makes sense to supply the air where the fresh air demand is high and the odor and humidity load is relatively small, i.e. in bedrooms and living rooms.
Through open doors or (acoustically insulated) air supply openings, the air then ends up e.g. in a
corridor and from there out into the rooms with the most contaminants, namely in the bathroom/WC
and kitchen, from where it is ultimately extracted (Fig. 7.21).
Extract air

Supply air
Through-flow zone

Bedroom
Living room

Corridor

Bathroom
Toilet
Kitchen

Fig. 7.21: Principle of cascades for mechanical ventilation [7.12]

7.7

Passive Cooling by Night-time Ventilation

The air exchange between the interior and exterior can also be applied to cool the building in
summertime. Passive cooling by night-time ventilation takes advantage of the outside night airtemperature being mostly below the room temperature, which is generally the case in Switzerland
also in summer. If a window is opened in a building at night, then the cool air flowing in can reduce
the temperature of the interior building materials so that on the following day a heat sink is available to absorb the thermal load (Fig. 7.22). The windows can be opened and closed either manually
or with automatically functioning mechanisms ( rain- and wind sensors). Passive cooling by
night-time ventilation is the simplest method of building cooling and requires practically no auxiliary
energy! The climatic potential for this technology is considerable in a large part of Europe [7.13],
especially in central, eastern and northern Europe.

Internal gains

Solar
gains

Fig. 7.22: Principle of passive cooling by night-time ventilation: charging (warming up) and discharging (cooling down) the massive structural building elements with heat in day/night cycles

To reach a high level of thermal comfort in a non-airconditioned building, it is important, along with
sufficiently cool air at night [7.14]:
minimize the thermal loads
sufficiently large air exchange ( cross ventilation)
enough (activatable) thermal mass in the room interior (e.g. concrete ceiling)

Air Exchange

142

This means that the external loads (size and orientation of the glazed areas, solar protection devices) as well as the internal loads (appliances, lighting) must be kept as small as possible. The
larger the air exchange during the night the better the heat can be removed. With cross ventilation,
the (window) openings are here placed on opposite sides of the room; a higher air exchange can
practically always be achieved than with a single-sided ventilation. Driving forces are temperature
differences as well as wind-induced pressure differences. The interior building elements must possess a high heat capacity to produce a satisfactory effect. Concrete ceilings are suitable for this but
also massive walls and floor construction. Thermal resistances between the room air and storage
elements have an adverse effect, such as e.g. suspended ceilings, subfloors or carpets. In todays
office building construction the concrete ceiling is often the most important storage element.
Figure 7.23 shows a section through an office building with an atrium in which the cold external air
flows through open windows, ideally flowing primarily as as a jet along the exposed concrete ceiling (bottom hung window) and in doing so absorbing the heat, and subsequently leaving the room
through a second opening on the opposite side of the room (cp. also Fig. 7.22). In the middle of the
atrium the warm air can rise and escape to the outside through openings in the roof.
Office

Atrium

Office

Fig. 7.23: Passive cooling by night-time ventilation: airflow in an office building during the night
[7.15]

Problems
Problem 1: Air exchange in an assembly hall
An assembly hall with a volume of 800 m3 offers space for a maximum of 100 people.
a) How great must the air exchange with the exterior be, in order to guarantee the air quality in the
interior (carbon dioxide and odor)? Calculate the air exchange n assuming that a person requires 15 m3/h of fresh air.
b) The space has two opposite walls each with windows with a total joint length of 80 m each. All
the air leakage coefficients are 0.05 m3/(hmPa2/3). How great is the air exchange that arises
with a pressure difference of 10 Pa over each of the joints of both walls (high- and low pressure
respectively)?
Problem 2: Wind pressure and air exchange
A floor plan for an apartment with a volume V = 200 m3 is given (drawing). The facades 1 and 2
have windows with the following joint characteristics:

Air Exchange

a1 = 0.2 m3/(h m Pa2/3) and l1 = 36 m

a2 = 0.2 m3/(h m Pa2/3) and l2 = 18 m

143

Facade 1 is blown on by wind with a velocity of v0 = 5 m/s ( = 1.2 kg/m3). The pressure coefficients are Cp1 = 0.8 and Cp2 = -0.4. Calculate:
a) the pressures p1 and p2 on the facade (Note: p1,2 = C p1,2

1
2
v0 )
2

b) the relative interior pressure pi (Note: It can be assumed that all the interior apartment doors are
open so that the same pressure prevails in the entire apartment).
c) the air flow rate V through the apartment.
d) the air change rate n between the interior and exterior.
e) the ventilation heat flow Q (heat loss) with a temperature difference of 30 K between the interior and exterior ; ca a = 1200 J m 3 K
1

Problem 3: Wind pressure on a living room glazing


How great is the wind pressure (assume stagnation pressure) and the force on a living room glazing (3.62 m x 2.48 m) with each of the following wind velocities?
v1 = 50 km/h

v2 = 100 km/h

v3 = 150 km/h

( = 1.2 kg/m3)

Problem 4: Stack effect


In a 12 m high building a large window is opened below (z = 0 m), otherwise all the openings are
closed. The interior temperature is 22C, the exterior temperature and atmospheric air pressure
are e = 0C and pe = 95'000 Pa respectively. The roof is finished on the inside with planking without a moisture or air barrier and is not airtight (a = 0.2 m3/(h m Pa2/3), joint length l = 120 m).
a)

Where is the neutral zone located? Draw qualitatively the interior/exterior pressure profile.

b)

What interior/exterior pressure difference results on the roof?

c)

How great is the air flow rate through the planking?

Air Exchange

144

d)

With this, how much water vapor is transported per hour if the room humidity is WD = 10 g/m3
(corresponding to 50 % relative humidity in the interior)?

Problem 5: Characteristic air-tightness values of a building envelope


In a single-family house (interior volume V = 400 m3, envelope area Ae = 400 m2) the values D =
100 m3/(h Pam) and m = 0.6, respectively, have been experimentally determined for the function
V = D Dp m .
a.) Calculate the n50- and va4-values.
b.) What relative pressure (sign?) develops in the building, if without additional openings an airhandling unit with an air flow rate of 100 m3/h is put into operation?

Literature
[7.1]
[7.2]
[7.3]
[7.4]

[7.5]
[7.6]

Moser A., Dorer V., Grundlagen der Raumluftstrmung, Bundesamt fr Energiewirtschaft


(BEW) und Verband Schweizerischer Heizungs- und Lftungsfirmen (VSHL), 1994
SIA 180, Wrme- und Feuchteschutz im Hochbau, Schweizerischer Ingenieur- und Archtekten-Verein, Zrich, 1999
Moor H., Physikalische Grundlagen der Gebudeaerodynamik im Hinblick auf die Berechnung des Luftaustausches, Eidgenssische Materialprfungs- und Forschungsanstalt, 1987
Manz H., Huber H., Helfenfinger D., Lftungstechnische und energetische Eigenschaften
von Einzelraumlftungsgerten mit Wrmerckgewinnung, 11. Schweiz. Status- Seminar
Energie und Umweltforschung im Hochbau, ETH Zrich, 2000
Meteonorm 6.0 (Edition 2007), Datenbanksoftware, Meteotest, Bern;
http://www.meteonorm.com
Zrcher C., Frank T., Bauphysik: Bau und Energie, vdf, Zrich, 2004

[7.7]

EN 12207, Fenster und Tren Luftdurchlssigkeit - Klassifizierung, Europisches Komitee


fr Normung, Brssel, 1999

[7.8]

Bundesamt fr Gesundheit; http://www.bag.admin.ch/themen/strahlung, 2008

[7.9]

SN EN 13829, Wrmetechnisches Verhalten von Gebuden Bestimmung der Luftdurchlssigkeit von Gebuden Differenzdruckverfahren, Europisches Komitee fr Normung,
Brssel, 2000

[7.10] Passivhaus Institut, D-64283 Darmstadt; http://www.passiv.de, 2008


[7.11] Recknagel/Sprenger/Schramek, Taschenbuch fr Heizung + Klimatechnik, R. Oldenburg
Verlag, Mnchen, 1997
[7.12] Huber H., Komfortlftung Projektierung von einfachen Lftungsanlagen im Wohnbereich,
Faktor Verlag, Zrich, 2004
[7.13] Artmann N., Manz H., Heiselberg P., Climatic potential for passive cooling of buildings by
night-time ventilation in Europe, Applied Energy, Vol. 84, 2007, 187-201
[7.14] Artmann N., Manz H., Heiselberg P., Parameter study on performance of building cooling
by night-time ventilation, Renewable Energy (available online April 7, 2008)
[7.15] Bob Gysin & Partner BGP Architekten, Eawag Forum Chriesbach ein nachhaltiger Neubau, Zrich, 2006

Chapter 8
Non-Steady-State Behavior of a Room

i
Pint
GI(t)

e
H[i(t)- e(t)]

Phc

System

Non-Steady-State Behavior of a Room

146

Non-Steady-State Behavior of a Room

Buildings are subjected to variable boundary conditions with respect to time: given externally by
the weather and internally by the use. Accordingly the temperatures of the building elements and
the interior air fluctuate. In this chapter it will be shown which parameters determine the nonsteady-state (= dynamic) thermal behavior of the interior and the influence of these parameters will
be discussed. The fundamental strategy that leads to the minimization of energy needs for heating
and cooling will be introduced and the consequences for design and construction will be drawn-out.
In the planning practice simulation programs are increasingly used to investigate the non-steadystate thermal behavior of buildings as well as to optimize the comfort and minimize the energy requirements. It will be illustrated with examples how the thermal comfort in an office building in
summer can be analyzed with these tools.

8.1

Energy flows in a room

A room or an entire building represents an energy system (Fig. 8.1), in which different energy flows
occur. These flows transfer energy in or out of the room. Thereby the stored heat in the room and
consequently also its temperature is changed. The following energy flows occur:
-

transmission heat flow


ventilative heat flow
solar radiation
heat absorption and heat release of building elements
internal heat sources
heating and cooling power

i
Pint

e
H[i(t)- e(t)]

Phc

GI(t)

Fig. 8.1: Energy flows in a room


System boundary

Corresponding to the temperature difference between the interior and exterior, heat flows through
all elements of the building envelope by transmission. The heat flow is greater the larger the thermal transmittance coefficient U and area A of the individual building elements are.
k tot

Transmission heat flow:

Q T = ( i e ) Ak U k

(W)

k =1

i
e
Ak
Uk

interior temperature
exterior temperature
area of k-th building element (wall, window, roof, etc.)
thermal transmittance of k-th building element

C
C
m2
W/m2K

The greater the air exchange and the temperature difference between the interior and exterior are,
the greater is the heat flow due to ventilation.

Non-Steady-State Behavior of a Room

Ventilative heat flow:

QV =

n
V a ca ( i e )
3600

147

(W)

n
V
a

air change rate


volume of room
density of air ( a 1.2 kg/m3)

m3

1/h

ca

specific heat of air ( c a 1005 J/kgK)

J/kgK

kg/m3

The energy exchange with the environment takes places at the exterior surface of the building,
which is why the parameters below relate to the building envelope area Ae [8.1] (Note: In contrast
to this the energy consumption of a building is in general related to the floor area (cp. Chapter 9).
Ventilative and transmission heat losses can be described together with a loss coefficient (without
thermal bridges).
Mean loss coefficient:
Ae

H=

n V a ca
1 ktot
Ak U k +

Ae k =1
3600

(W/m2K)
m2

total building envelope area

Solar energy can enter the building interior through transparent building elements. Decisive is the
total solar energy transmittance (section 6.1) of the different elements as well as their respective
areas.
G=

Mean total solar energy transmittance:


gk

1 ktot
g k Ak
Ae k =1

total solar energy transmittance of k-th building element

(-)
-

Heat can be stored in the interior building elements such as ceilings, walls and floors and later released again. The specific heat per area is:
Ck = k c k d k

k
ck
dk

(J/m2K)

density of k-th building element


specific heat of k-th building element
effective thickness for heat storage of k-th building element
(see section 5.2.3)

kg/m3
J/kgK
m

The heat capacity of individual building elements can be added and related to the building envelope area.
Mean specific heat:

C=

1 ktot
Ak k c k d k
Ae k =1

(J/m2K)

The thermal power that is given off by people and appliances as well as the heating and cooling
power can likewise be related to the building envelope area.
Interior heat sources:

P' int =

Heating/cooling power:

P' hc =

PPers + PApp
Ae
Phc
Ae

(W/m2)

(W/m2)

Non-Steady-State Behavior of a Room

148

8.2

Energy Balance in a Room

The equation for the conservation of energy can be applied to the room (Fig. 8.1). According to this
equation the difference in the energy flowing in and flowing out of the room equals the change of
stored energy in the room:
G I (t ) + P 'int +P 'hc H [ i (t ) e (t )] = C
in

out

d i
dt

(W/m2)

= increase

solar irradiance

W/m2

time

If one assumes that the room is neither heated nor cooled (Phc = 0), and also no heat is given off
by people and appliances (Pint = 0), then the natural room temperature or free-running temperature
is reached. The energy balance equation can be written as follows [8.1]:

i (t ) +

G
C d i

= e (t ) + I (t )
H dt
H

(C)

The time-dependent room temperature i is on the left side of the equation, and its derivative (=
change) respectively. On the right side of the equation is the time-dependent weather, namely the
external temperature e and the solar irradiance I. Interestingly the characteristics of the room are
described by only two characteristic values.
Time constant:

C
H

(s or h)

Gain/loss-ratio:

G
H

(m2K/W)

8.3

Time Constant and Gain/Loss-Ratio

As a first example, the cooling off of a building is considered to illustrate the time constant. Under
the assumption that neither solar gains (G = 0) nor internal gains (Pint = 0) occur and it is neither
heated nor cooled (Phc = 0), the room temperature with respect to time i (t) can be found analytically from the solution to the following equation:

i (t ) e + t

d i
=0
dt

(C)

The external temperature e is assumed to be constant. An exponential function of the form

i ( ) = e + [ i ( = 0 ) e ] e

(C)

i (t = 0 )

room temperature at time t = 0

time constant

s or h

describes the solution to this equation. Figure 8.2 shows the influence of the time constant on the
cooling down process. At time t = 0 the room temperature is at 20C and the external temperature
is constant at 0C. The better the heat storage (C large) is, but also the better insulated and airtight (H small) a building is, the larger is its time constant. Also the building geometry is of importance: the larger and more compact a building is, the smaller is its envelope area per volume,
and the smaller is the thermal loss and the slower is the cooling down. The time constant thereby
designates the time with which the temperature difference between the interior and exterior sinks
to 1/e (= 36.8 %).

Non-Steady-State Behavior of a Room

Internal temperature (C)

20

149

well insulated, airtight, large building with a


high heat capacity

= 500 h

15

= 100 h

i (t ) = i (t = 0 ) e

10

= 20 h
0

72

48

24

poorly insulated, non-airtight, small building


with a low heat capacity
96

120

Time t (h)

Fig. 8.2: Cooling down of buildings with different time constants


The second example illustrates the significance of the time constant , as well as the gain/loss-ratio
. Figure 8.3 shows the external temperature and the solar irradiance on a south facade in May in
Zurich. Assumed is a building with no interior gains (Pint = 0) and which is neither heated nor
cooled (Phc = 0):

i (t ) + t

d i
= e (t ) + I (t )
dt

(C)

30

500

External temperature (C)

In the Figures 8.4 und 8.5 the numerical solutions to this equation are shown for the weather data
from Fig. 8.3 with different values of and ..

20

400

0
200
-10
100

-20
-30
0

10

15
Time t (d)

20

25

Fig. 8.3: Weather data May 2005 (Zurich-SMA)

30

Solar irradiance

300

south facade I (W/m2)

10

Non-Steady-State Behavior of a Room

150
35

= 20 h

30

Room temperature (C)

= 0.1 m2K/W

25

= 500 h

Comfort range

20
15
10
0

10

15

20

25

30

Time t (d)

Fig. 8.4: Influence of time constant on the free-running temperature of the room

= 100 h

30

Room temperature (C)

35

= 0.15 m2K/W

25

Comfort range
20

= 0.05 m2K/W
15
10

10

15

20

25

30

Time t (d)

Room temperature

Fig. 8.5: Influence of the gain/loss-ratio on the free-running temperature of the room

Too warm

Comfort range

Too cold

Fig. 8.6: Influence of and on the room temperature


In general the time constant represents a measure for the thermal inertia of a building or room,
i.e. the ability to remain unresponsive. The room is less responsive, the greater its storage ability,
i.e. its specific heat per area (C) is, and respectively the smaller its thermal losses (H) are. The
time constant increases, and respectively decreases, the amplitude of the room temperature fluctuations (Fig. 8.6). The gain/loss-ratio shifts the free-running temperature, and also influences the
amplitude. High values of produce high room temperatures and large amplitudes. Buildings with

Non-Steady-State Behavior of a Room

151

large time constants are very fault tolerant. If for example the heating system fails temporarily, it
will have almost no effect on the room temperature.
Figure 8.7 schematically shows the external temperature over the course of a year and the comfort
range. In summer a somewhat higher temperature is preferred given that the people are then
mostly lightly clothed. To design buildings that need little operating energy, the values of and
should be selected such that the free-running temperature (including the heat given off from people
and appliances) remains as long as possible in the comfort range. No or virtually no heating and
cooling is then required.

Temperature

Poor building

Ideal building
Comfort range
External air

Winter

Summer

Winter

Fig. 8.7: Free-running temperature of buildings and comfort range


The technical means to influence the two parameters and are:
Window areas, solar transmittances, orientations, solar protection devices (variable)

Gain/loss-ratio:

Time constant:

G
H

C
=
H

U-values, areas, air exchange possibly with


heat recovery, nigh-time ventilation (variable)

Internal surface areas, heat capacity of materials,


layer thickness
U-values, areas, air exchange possibly with
heat recovery, nigh-time ventilation (variable)

In the selection of the two parameters and it should be noted that


- the time constant should be large to keep the room temperature fluctuations small
- a variable gain/loss-ratio makes it possible to influence the level of the free-running temperature and to keep it in (or nearby) the comfort range ( summer/winter)

Therefore, it is beneficial for the thermal behavior of a building to have:


- a small loss coefficient (H ):
- a good insulation, which is free of thermal bridges
- small (thermally relevant) air exchange:
- air-tight building envelope

Non-Steady-State Behavior of a Room

152

- heat recovery in the air-handling unit


- large heat storage capacity (C )
- variable solar gain: variable solar protection devices (G )
The building geometry is also relevant: Buildings with a small ratio of the envelope area to volume
i.e., large compact buildings tend to exhibit higher time constants.

8.4

Building Simulation

The earlier considerations (section 8.2 und 8.3) are helpful in order to develop an understanding
for the non-steady-state behavior of a building. As already mentioned, the parameters and are
not constant for actual buildings because solar shading devices and windows can be opened or
closed whereby the characteristics of a building are considerably changed. Additionally, internal
sources are time-dependent and the solar irradiance varies with time, depending however also on
the orientation of the windows. Often the layer sequence of a construction is relevant to how well
the heat can be stored. For precise questions regarding non-steady-state behavior of a building it
is for these reasons common in practice to use the method of numerical building simulation [8.2]
instead of an analytical or semi-analytical approach.
Building simulation can be applied particularly:
- to optimize the thermal comfort in summer and minimize the cooling load respectively, especially for buildings with large glazed areas and relatively high internal loads (commercial buildings)
- to size cooling units (required maximum output) as well as to investigate control strategies
- to design low energy buildings: optimization of solar gains, sizing of solar collectors, etc.
The method of building simulation is based on a thermodynamic, as well as a fluid-dynamic, network model (Fig. 8.8). The strength of the method lies in that all the relevant energy flows in a
building building structure and HVAC systems can be simulated at the same time. The external boundary conditions are given by the weather (air temperature, solar irradiance, etc.) and the
internal boundary conditions through the use (heat given off from people and appliances). The nonsteady-state heat transfer through the building envelope, but also the heat storage processes in
the interior building elements, are taken into account. In the interior rooms it is often differentiated
between convective and radiative heat transport. The overall energy conservation equation in the
programs is typically solved in intervals of one hour. Many programs also allow the modeling of
multiple zones.
e
e

IS
IN

i,1
e
i,2

i,3

e
g

Fig. 8.8: Network model for building simulation

Non-Steady-State Behavior of a Room

8.5

153

Building Simulation Example: Office Room in Summer

External
temperature (C)

Solar irradiance (W/m )

In our climate particularly office buildings tend to overheat in summer because the internal loads
(people and appliances) are generally higher than in residential buildings. In addition, modern office buildings often are extensively glazed, that can also depending on the solar protection devices and facade orientations lead to considerable solar loads. West-orientated rooms are often
at greatest risk, because in the late afternoon high solar irradiances and high external air temperatures occur simultaneously (Fig. 8.9), and at a time of day when the room can already be considerably heated up.

Time (h)

Fig. 8.9: Typical daily cycles of solar irradiances and external air temperature [8.3]

To illustrate the building simulation method [8.4] a west-orientated office room is considered with a
4 m 5 m floor area, a room height of 2.6 m and a percentage of glass to facade area of 40% (0.4
2.6 m 4 m = 4.16 m2 glass surface). It is assumed that the energy exchange between the room
and the environment takes place solely over the facade (= adiabatic boundaries at the neighboring
zones in the building). Windows and solar shading devices exhibit the following characteristics:
- thermal transmittance Uw = 1.1 W/m2K
- total solar energy transmittance without shading, gwithout = 0.6
- total solar energy transmittance with shading gwith = 0.15
- solar shading device closed, if I > 300 W/m2
The air exchange rate will be different according to the time of day, whereby a higher air exchange
occurs at night for passive cooling (section 7.7):
- day:

n = 2 1/h

- without night-time ventilation:

n = 0.5 1/h

- with night-time ventilation:

n = 6 1/h

Non-Steady-State Behavior of a Room

154

Ceiling, floor and wall of the room are constructed according to Fig. 8.10.

internal wall

external wall

ceiling and floor

carpet 0.5 cm
screed 8 cm
impact sound insulation 4 cm
concrete 18 cm

gypsum board 2.5 cm


mineral wool 7 cm
gypsum board 2.5 cm

interior plaster 1.5 cm


sand-lime brick 15 cm
insulation 15 cm
exterior plaster 2 cm

Fig. 8.10: Construction of ceiling, floor and wall


The internal heat loads are made up of the heat given off from people, appliances and lights and
have a typical output in the course of the day for office use (Fig. 8.11).

30

Last (W/m )

25
20
Appliances
Gerte
Lighting
Beleuchtung
People
Personen

15
10
5
0
7

10 11 12 13 14 15 16 17 18 19 20
Time
Zeit (h)
(h)

Fig. 8.11: Course of internal loads in the day


With the help of building simulation it can now be investigated how both measures, solar shading
and passive cooling by night-time ventilation, have an effect on the summery comfort. Figure 8.12
shows the course of the external temperature and solar irradiance on a west-facade in Zurich during a week in June. This course comes from a semi-synthetic DRY-weather data set (DRY = Design Reference Year) that corresponds to a typical year. In the simulation the time period from
May 1 to September 30 is used.
The effectiveness of the solar shading and night-time ventilation is apparent in Figure 8.12. The
comfort level reached is indicated through the operative room temperature (section 3.3). If neither
solar shading nor night-time ventilation is deployed, then the operative room temperature stays in a
range of about 33C to 40C! If either the solar shading or the night-time cooling is used then the
average room temperature can be reduced by about 8 K. On sunny days the solar shading is more
effective and on cloudy days the night-time ventilation is. The combination of both measures leads
to a further reduction of the operative temperature by about 4 K.

Non-Steady-State Behavior of a Room

155

32

800

Temperature
(C)
Temperatur (C)

Temperature

26

600
Solarstrahlung

20
14

400
200

8
17.06.

18.06.

19.06.

20.06.

21.06.

22.06.

23.06.

0
24.06.

42
38
without shading / without night ventilation

34
30

with shading / without night time ventilation

26

without shading / with night-time


ventilation
with shading / with night-time ventilation

22
18
17.06.

18.06.

19.06.

20.06.

21.06.

22.06.

23.06.

24.06.

Fig. 8.12: Course of exterior climate (above) and operative temperature of the office room (below)
during a week in June in Zurich
40
Operative room temperature (C)

Operative Raumtemperatur
room temperature (C)
Operative
(C)

Zurich DRY

Solarstrahlung (W/m2)
2
Solar irradiance (W/m )

Figure 8.13 shows the cumulative frequency distribution of the operative room temperature with the
different measures. In this plot it is apparent what comfort level is reached at what amount of the
time. If both solar shading and night cooling are used, the operative room temperature during the
entire summer will remain below 26.5C (= good comfort level) for the subject office room.

without shading / without night-time ventilation

too warm
35

without shading / with night-time ventilation


with shading / without night-time ventilation

30

with shading / with night-time ventilation

25
good comfort
20

0.2

0.4

0.6

0.8

Fraction of time (-)

Fig. 8.13: Cumulative frequency distribution of the operative room temperature


(May 1 September 30)

Non-Steady-State Behavior of a Room

156

The strength of the method of building simulation lies particularly in that parameter changes can be
quickly carried out and their impact on the level of comfort analyzed. Already in the planning phase
it can be shown which comfort level can be reached with a given configuration of glazings, solar
shading devices, building elements (walls, ceilings and floors), air exchange and internal loads.
Generally the following strategies help to achieve a high comfort level in summer without active
cooling:
- minimize internal loads: energy efficient appliances and lighting
- minimize solar gains: areas and orientation of windows, effective (g < 0.15) solar shading devices
- good use of daylight: minimize electrical lighting (better luminous efficiency with sunlight!)
- large storage capacity of interior building elements: ceilings, walls and floors (Attention: thermal
decoupling)
- passive cooling with night-time ventilation: storage capacity of interior building elements is important!

8.6

Structural Consequences

In this section the parameters that determine the thermal dynamic behavior of an interior room or
building (H, C and G and respectively and ) have been presented. The influence of these parameters on the free-running temperature of the building has been illustrated. The goal is to keep
as long as possible the free-running temperature in the comfort range! Then a building can be realized that hardly needs energy for heating and cooling. This results in consequences for design and
construction:
- loss coefficient H as small as possible: good insulation, high airtightness, mechanical ventilation
with heat recovery, only ventilate as much as necessary
- heat storage capacity C as large as possible: interior building elements
- variable solar energy transmittance G: variable (external) shading devices
The non-steady-state behavior of buildings can be significant particularly in summer. Building simulation presents a valuable tool for analyzing problems of overheating in summer.

Problems
Problem 1: Cooling Down of a Single-Family House
The heating system of a single-family house heated to 20C suddenly fails in winter. Calculate, under the assumption that there are no heat sources in the interior and no solar gains, the interior
temperature after one hour, one day and one week. The external temperature of the environment
is constant at 4C. The ratio between the heat capacity and heat losses of the building is C/H =
200 h.

Problem 2: Warming of a Cold Storage House


A windowless, cubic, empty cold storage house out of concrete with an edge length of 10 m is in
contact on all sides with the exterior climate and has a mechanical air exchange of n = 0.5 h-1 with
a cold recovery (heat exchanger, h = 0.8; air: a = 1.2 kg/m3, ca = 1005 J/kgK). The air exchange

Non-Steady-State Behavior of a Room

157

due to infiltration is negligible. The building envelope is identically constructed on all six sides: 15
cm concrete on the inside (C = 2.5 W/mK, C = 2400 kg/m3, cC = 1000 J/kgK) and 20 cm insulation
on the outside (Ins = 0.04 W/mK). The interior room is cooled to 5C, the exterior is at a constant
temperature of 20C. Suddenly the cooling system fails.
a.)

b.)
c.)
d.)
e.)
f.)
g.)
h.)

Calculate the heat capacity C of the building. It can be assumed that the entire concrete
thickness is effective in heat storage. The storage capacity of the insulation can be neglected.
Calculate the loss coefficient H of the building.
Write the balance of energy flows.
Formulate from c.) the emerging differential equations using the time constant .
Determine the analytical solutions of the differential equations.
How large is the time constant of the building?
What temperature can be expected after 24 h, 72 h, 240 h and 720 h respectively in the interior?
Graphically illustrate the interior as well as the exterior temperature in the time interval 0 t
1200 h.

Literature
[8.1]

Keller B., Bauphysik: Die Energetik des Gebudes, Vorlesungsskript ETH, Zrich, 2006

[8.2]

Clarke J.A., Energy Simulation in Building Design, Butterworth-Heinemann, London, 2001

[8.3]
[8.4]

Zrcher C., Frank T., Bauphysik: Bau und Energie, vdf, Zrich, 2004
HELIOS, Gebudesimulationsprogramm, Eidgenssische Materialprfungs- und Forschungsanstalt, Dbendorf, 2008

Chapter 9
Energy and Sustainability

Image: The Earth at night (a composite assembled from satellite images from NASA)

Energy and Sustainability

160

Energy and Sustainability

Without the large use of energy human development would have taken an entirely different course
in the last 200 years. The cheap availability of energy significantly shaped the development of the
industry and service sectors as well as our society. Many technical achievements that we use today are based on the use of energy. Energy consumption and economic performance appears to
be closely dependent on one another, whereby very large differences between different countries
exist (Fig. 9.1).

Katar

Energy consumption per capita


(kg oil equivalent/person year)

104

UAE

Bahrain
USA
Russia

Germany
Japan
Switzerland
UK

Usbekistan

South Africa

1000

Indonesia
Nigeria

China

India
Congo
Ethiopia
Eritrea

100

Mexico
Brazil

Columbia

Pakistan
Senegal

Source: World Resource Institute &


CIA World Factbook, 2007

Bangladesh

1000
104
Gross national product per capita
(US $/person year)

105

Fig. 9.1: Energy consumption and gross national product per capita in different countries
The use of energy is however increasingly linked with problems like environmental pollution, climate change and shortage of resources. This set of problems will be briefly outlined in the following.
Energy is needed in transportation, buildings, industry and agriculture. Nearly half of the energy
demand in Switzerland arises in buildings: heating, cooling, hot water, ventilation, lighting and appliances. The building envelope design especially determines, quite significantly, the emerging energy demand. The expenditure of energy must be considered during the full life cycle of a building:
construction, operation, refurbishment, and lastly demolition (embodied energy). All of these aspects will be highlighted and the strategies for sustainable buildings depicted. Thereby the reduction of energy required for operation has a high priority.

9.1

Energy and Sustainability Challenges

The world population has dramatically increased in the last century (Fig. 9.2). Today (2014) 7.3 billion people live on our planet. This corresponds to a growth by a factor more than four since the
year 1900. According to a prediction by the UN the world population will increase to nearly 10 billion by the year 2050, although the yearly growth is declining since about 1990 (2014: growth
about 80 million).

Energy and Sustainability

161

7000
2000: 6.07 billion

6000
Population (millions)

5000
4000

+ 268%

3000
2000
1900: 1.65 billion
1000

+ 68%

1800: 0.98 billion


1700: 0.6 billion

0
-1000

-500

500
Year

+ 63%

1000

1500

Source: U.S. Census Bureau, Population Division, 2007

2000

Fig. 9.2: Development of the world population

250
constant
consumption rate

200

1 % increase
per year

150
100
50
0
2000

2050

2100

2150

2200

380
360

Keeling et al., 2005


(Mauna Loa, atmosphere)

340

5 % increase
per year

Atmospheric CO concentration (ppm)

Remaining resources: remaining time for


use at todays consumption rate (Years)

In the beginning of the 1970s the general public became aware of the finiteness of the Earth and
the ramifications of unregulated growth. This was, among other things, a consequence of the Club
of Rome report The Limits to Growth [9.1]. In the earths system both the sources, i.e. the natural
resources of materials and fossil fuels, and the sinks, i.e. the capacity of the environment to absorb
all kinds of pollution as well as waste heat, are limited. Figure 9.3 (left) shows the estimated future
depletion of the worlds natural gas reserves under the assumption that the present day supply can
cover todays consumption for 240 years. This means that three times more reserves exist than are
known today. As an example for the anthropogenic burden on sinks, Figure 9.3 (right) shows the
increase of the carbon-dioxide content in the atmosphere. This increase is a result of burning fossil
fuels and the slash-and-burn of large forest areas. As has long been known, the increase of carbon
dioxide in the atmosphere leads to the decrease of long-wave thermal radiation of the earth to outer space and thereby to a warming of the earths surface with partially unpredictable consequences
for people, animals and vegetation. The burden on sinks is largely seen as a greater threat than
the depletion of resources.

2250

320
Etheridge et al., 1998
(ice cores)

300
280
1830 1860 1890 1920 1950 1980 2010

Fig. 9.3: Depletion of global natural gas reserves [9.2](left) and increase of carbon dioxide in the
atmosphere (right)

Energy and Sustainability

162

Calculations from environmental economists have shown that the total human activities today exceed the long-term carrying capacity of the earth by roughly a fourth (Fig. 9.4). That is to say, we
actually need 1 Earths to be able to cover our current needs. This means that we use up resources faster than they are regenerated. In this context the term ecological footprint is often used.
Like the per capita energy consumption, the ecological footprint per capita also varies very strongly
according to country.
1.4
ecological footprint of humanity

1.2

Number of Earths

1
carrying capacity
of the Earth

0.8
0.6
0.4

0.2
0
1960

1970

1980

1990

2000

Fig. 9.4: Development of the ecological footprint of humanity compared to the long-term carrying
capacity of the Earth [9.3]
In addition to the mentioned global limitations of resources and absorption capacity of the environment, it is also essential to contain local pollution. Especially in air of urban areas, mainly as a consequence of combustion, considerable concentrations of particulate matter, nitrogen oxide (NOx),
volatile organic compounds (VOC), sulfur dioxide (SO2) and ozone (O3) can arise that lead to respiratory diseases.

Fig. 9.5: Development of energy consumption in Switzerland (1910-2006)[9.4]

Energy and Sustainability

163

In the course of the 20th century the energy consumption in Switzerland (Fig. 9.5) increased by a
factor of about eight! A notable growth in electricity, natural gas and petroleum took place in the
last 30 years. The consumption of oil has slightly declined since the 1970s. The renewable energy
(wood, solar, geothermal, wind) still plays a marginal roll at this time. Overall consumption today
appears to have somewhat stabilized but at a high level.
A totally different picture arises worldwide (Fig. 9.6). The international energy agency expects that
from the years 2004 to 2030 the global energy demand will increase by about a half! Considering
the population growth (Fig. 9.2) and the fact that the per capita consumption in many developing
and emerging countries today still lies below Europe or North America by a factor of five to ten (i.e.,
that in many countries a huge pent-up demand for energy exists), this is not surprising.

Primary energy (Mtoe)

Reference Scenario

Fig. 9.6: Development of global energy demand [9.5]


The use of fossil fuels (natural gas, oil, coal) gives rise to an immense release of carbon-dioxide.
The worldwide CO2-emissions in the year 2004 amounted to about 26 gigatons (Giga = 109). According to the International Energy Agency [9.5] this value will increase by the year 2030 from 34
Gigatons per year (alternative policy scenario) to 40 gigatons per year (reference scenario). According to [9.5], a stabilization of emissions will be possible in 2030, at the earliest, particularly if
energy efficiency and renewable energy are pushed further.

2 % Others

28 % Mobility (domestic)
35.1 % Space heating

8.9 % Mechanical power,


other processes
13.0 % Process heat

5.5 % Hot water


2.7 % Ventilation & air conditioning
3.4 % Lighting
1.4 % Information & communication

Fig. 9.7: Energy use in Switzerland [9.6]

Energy and Sustainability

164

The distribution of the Swiss energy requirements according to end-use shows (Fig. 9.7) that just
under half of the requirements arise in buildings. The most important designated use is space heating (35.1%), followed by hot water (5.5%), ventilation and air conditioning (2.7%) and lighting
(3.4%). The sum gives that about 47% of the energy requirements in Switzerland occur in buildings.
Also in many other industrialized countries buildings are responsible for a large portion of the total
energy needs. For example, in the USA 39% of the primary energy is used in buildings, 29% in
transportation and 32% in industry [9.7].
If different energy sources are assessed and compared with one another then the entire chain of
energy conversion must be considered. So, for example, the crude oil from boreholes (primary energy) is converted in the refinery into fuel oil (secondary energy), afterwards it is delivered to the
consumer (end energy) and in the end converted into heat (useful energy). The definitions of terms
used in the energy business are given in Table 9.1.

Definition

Example

primary energy

unprocessed energy

crude oil, coal, uranium, hydro energy, solar


energy

secondary energy

(at least once) processed energy

heating oil, fuel rods for nuclear power plants

end energy

energy delivered to the consumer

heating oil (delivered), electricity

useful energy

effectively used energy

heat, light, process energy, kinetic energy

Tab. 9.1: Definition of terms from the energy business


Buildings are not only important with respect to energy, but also regarding the flow of materials,
because about half of the mass flow that our society generates is incurred in the construction sector (Fig. 9.8). From this point of view the building can be seen as a gigantic temporary store of materials. With renovation and demolition, waste is generated which is to date only partially recycled.
Therefore, also waste disposal requirements arise and with that a need for especially in densely
populated areas valuable land. With this the construction waste, depending on its composition,
can leach pollutants into the soil and ground water over time at the disposal site. For that reason
with the selection of materials not only the costs and appearance should be considered, but also
the contaminant potential, the recyclability, the embodied energy and the influence on the direct
energy consumption during operation.

Fig. 9.8: Material flows in the Swiss construction sector [9.8]

Energy and Sustainability

9.2

165

Heating Energy Demand

9.2.1 Balancing the Energy Flow in a Building


In this section it will be shown how the heating energy demand of a building can be calculated. In
general only steady-state models are used [9.9, 9.10]. Energy that flows in and out of a building is
balanced (Fig. 9.9) over a long time period, such as a month. The net heat loss corresponds to the
heating energy demand Qh that is required to maintain a constant temperature in the building. It
can be calculated as follows (nomenclature according to SIA 380/1):
Qh = QT + QV h (Qi + Qs )

QT
QV

h
Qi
Qs

(J)

transmission losses
ventilation losses
utilization factor
internal gains (people, appliances, lighting)
solar gains

J
J
J
J

Not all internal and solar gains are usable. The utilization factor corresponds to the portion of heat
gain that can be used for space heating and is dependent on the heating demand and the storage
capacity of the building, the control of the heating system (response time) as well as the users acceptance of temperature fluctuations in the interior. The utilization factor for heat gain h can be approximately determined from Figure 9.10 [9.9]. In this figure the utilization factor depends on the
gain/loss ratio; the ratio between the total heat gain (Qi + Qs) and the sum of heat losses due to
transmission and ventilation (QT + QV). Thus the utilization factor is high in winter and small in the
transition period. The time constant of the building is given by the ratio between the heat capacity
C and the loss coefficient H of the building, = C/H (compare chapter 8).

i
Qi
Qs

e
QT

Qh

system

Fig. 9.9: Energy flows in a building assuming steady-state conditions


The energy losses and respective gains can be calculated as follows (compare Chapter 8). Transmission losses (including thermal bridges) are considered according to [9.9].
jtot
itot
ktot

Transmission losses: QT = ( i e ) t Ak U k + l j j + zi i (J)


j =1
i =1
k =1

Energy and Sustainability

166

Time constant t

Utilization factor h (-)

0.9

50 h
100 h
200 h

0.8
0.7
0.6
0.5
0.4
0.3
0

0.5

1.5

2.5

Gain/loss ratio (-)

Fig. 9.10: Utilization factor for thermal gains in relation to the gain/loss ratio and the time constant
of the building [9.9]

i
e
t
Ak
Uk
lj

j
zi

interior temperature
exterior temperature
time period (e.g. 1 month)
area of k-th building element (wall, window, roof etc.)
thermal transmittance of k-th building element
length of j-th linear thermal bridge (e.g. roof/wall-interface)

C
C
s
m2
W/m2K
m

linear thermal transmittance of j-th thermal bridge


number of i-th point thermal bridges (e.g. pillar)
point thermal transmittance of i-th thermal bridge

W/mK

Ventilation losses:

QV =

n
V a ca ( i e ) t
3600

W/K

(J)

n
V
a

air exchange
volume of room
air density ( a 1.2 kg/m3)

1/h
m3
kg/m3

ca

specific heat of air ( ca 1005 J/kgK)

J/kgK

Solar heat gain:


Fk
gk
Gk

Qs =

ktot

Fk g k Ak Gk

(J)

k =1

reduction factor due to shading and frames of k-th building element


total solar energy transmittance of k-th building element
solar irradiation on the k-th building element

J/m2

Monthly mean values for exterior temperature and solar irradiation are available in the appendix for
a number of Swiss locations in different climate regions. Standard values for internal thermal
sources (people, appliances) for different uses (residence, office, indoor swimming pool, hospital
etc.) are available in [9.9].

Energy and Sustainability

167

9.2.2 Heating Degree Days


The transmission and ventilation losses during a certain time period are proportional to (i e)t.
One can assume that the heating system is only in operation if the time-averaged exterior air temperature e falls below a certain critical temperature, the so-called heating limit li. To simplify the
calculation of emerging heat losses during a heating period, the quantity HDD (Heating Degree
Days) is defined.

HD (Heating Days) denotes the number of days during which the time-averaged exterior temperature e falls below the heating limit li. HDD designates the area between the curves i and

e (t ) for e li (gray area in Fig. 9.11). The heating limit depends on the thermal quality of the
building. Well-insulated buildings exhibit low heating limit temperatures. The HDD- and HD-values
are available for different locations and different heating limit temperatures (cp. SIA 381/2 in the
Appendix).

Fig. 9.11: Daily mean values of exterior temperature e in the course of the year and heating degree days
The energy required to cover the transmission as well as the ventilation losses can be calculated
as follows:

QT = U A HDD 24 3600

(J )

QV = n V a ca HDD 24

(J )

The cooling degree days can also be calculated as with the heating, in order to calculate the cooling energy demand [9.11]. The accuracy is nevertheless not very high. Degree day methods provide however a very consolidated picture of the climate. The disadvantage lies in that a heating
and respectively a cooling limit must be assumed. In the current SIA 380/1 standard [9.9] the degree day method is not used for the heating demand calculation.

Energy and Sustainability

168

9.3

Protection Against Overheating in Summer

If a building is well designed it is possible in most cases in moderate central-European climates to


ensure comfortable conditions in the interior rooms in summer without active cooling. If the building
must nevertheless be cooled then the cooling load has to be minimized. Because comfort demands are increasing, buildings having large-area glazings are increasing (solar load), internal
loads are increasing (appliances, lighting) as well as gradual climate warming (Sect. 9.9), summer
heat protection is gaining, especially in commercial buildings, a lot more in importance. Within the
EU the floor area requiring air-conditioning has increased by a factor of 4 in the last 20 years (Fig.
9.12). The largest growth rate thereby is apparent in southern Europe (Fig. 9.13).

Fig. 9.12: Increase of cooled floor area in the 15 EU-countries [9.12]

Cooled floor area (m 2/inhabitant)

18
16
14
12
10

2000
2020

8
6
4
2

Be
D lgiu
en m
G ma
er rk
m
G any
re
ec
Sp e
Fr ain
an
Ire ce
la
n
Lu
xe It d
N mb aly
et o
he u
r la r g
n
Au ds
Po stri
rtu a
g
F
i a
U
ni S nla l
te w nd
d e
Ki de
ng n
do
EU m
-1
5

Fig. 9.13: Predicted change over time of cooled floor areas in the 15 EU-counties [9.12]

Energy and Sustainability

169

The measures to protect against overheating in summer can be grouped as follows:


- reduction of external loads (area and orientation of windows, shading devices, solar control
glazing, cp. Sect. 6.5)
- reduction of internal loads (lighting, appliances)
- increase of thermal storage capacity of the interior building elements (massive, thermally active)
- ventilation (passive cooling with nigh-time ventilation, cp. Sect. 7.7)
In order to allow for a reliable statement about the expected thermal comfort of a particular building
in summer, the process of thermal storage in the internal building elements has to be taken into
consideration. This can be accomplished with the help of building simulation programs (Sect. 8.5).

9.4

Renewable Energy

Energy is designated as renewable if it comes from a source that is inexhaustible from a human
scale. The sources for usable energy flow are thus:
- nuclear fusion in the sun and, therefore, solar irradiation: solar energy, hydro energy, wind energy, wave energy, bioenergy
- radioactive decay in the Earths interior: geothermal energy
- earths rotation: tidal energy
The so-called fossil fuels (crude oil, natural gas, coal) stem from energy from earlier solar irradiation. Because the time period of the formation (millions of years) in comparison to the time period
of utilization (a couple of hundred years) differs by many magnitudes these fuels are not considered renewable. For utilization in buildings, solar energy, geothermal energy and bioenergy are
most interesting as renewable energy.

Fig. 9.14: Fuel wood and carbon cycle


Wood is a renewable, carbon-neutral fuel. During the growth phase in the forest the same amount
of carbon-dioxide is absorbed as is released by burning (Fig. 9.14). Heating with wood is especially
prevalent in rural areas of Switzerland. The disadvantage is the emission of particulate matter that

Energy and Sustainability

170

result from burning. Improved combustion chambers and suitable electric filters can reduce the
emissions significantly. The latter are mainly used in large facilities.
In connection with buildings the use of near-surface geothermal energy is of interest. By means of
ground-coupled (earth-to-air) heat exchangers (Sect. 7.6), i.e. tubes laid approximately horizontal
in the earth, the supply air in a ventilation system can be pre-heated in winter and possibly precooled in summer.
Using ground loops, i.e. tubes placed in vertical boreholes in the soil for the heat exchange, the
heat is transported to the surface by means of a heat transfer fluid. Typically heat pumps are employed for heating buildings in such a system. An important characteristic value of heat pumps is
the coefficient of performance COP or, because the temperature boundary conditions change in
the course of a year, the seasonal energy efficiency ratio SEER. It gives the ratio of output heating
energy to input electrical energy over the year. A heat pump with a SEER of 3 produces based
on the intake of electrical energy triple the heat energy. The relatively high demand for electricity
can be considered a disadvantage of the heat pump.
In buildings solar thermal energy is today mainly used for providing hot water; used to a less extent
also for heating. For the temperature range from around 40C to 100C flat plate collectors are typical (Fig. 9.15). These solar collectors are in general covered with a glass pane to reduce the heat
loss. They have an absorber that is mostly selectively coated, a heat-transport fluid flowing through
the absorber to remove heat, and a heat insulating backing. With these solar collectors the direct
as well as the diffuse solar radiation can be utilized. The heat-transport fluid typically contains an
anti-freeze, to prevent freezing in winter.

beam

Tout

diffuse

heat-transport fluid

glass pane

thermal insulation
absorber

Tin

Fig. 9.15: Cross-section through a thermal solar collector (flat plate collector)
The efficiency of a solar collector can be derived from the law of conservation of energy applied to
the absorber:

t A I U A T = m c (Tout Tin )
gains

losses

(W)

useful heat

solar transmittance of glass pane


solar absorptance of absorber
absorber area

m2

Energy and Sustainability

171

I
U

solar irradiance
thermal loss coefficient

W/m2
W/m2K

temperature difference between absorber and environment


mass flow rate of heat-transport fluid
specific heat of the heat-transport fluid
input and output temperature

K
kg/s
J/kgK
C

m
c
Tin, Tout

The average temperature difference between the absorber and the exterior is approximately:

Tin + Tout
Te
2

The collector efficiency is given by the ratio of the useful heat, i.e. the convective heat flow due to
the heat-transport fluid, and the amount of irradiated solar energy:
(-)
The outcome of this is the collector efficiency:

= U

(-)

This is a linear equation: Collector efficiency h as a function of the boundary conditions T/I. The
product is also referred to as the optical efficiency. U corresponds to the slope of the straight
line. Different efficiencies are reached (Fig. 9.16) depending on the input temperature of the fluid,
the solar irradiation as well as the ambient temperature. In the simplified consideration above it
was assumed that ideal heat transfer exists between the absorber and the fluid (heat exchange
factor F 1).

Collector efficiency (-)

opticl losses
terml losses

0.5
U

useful et

U >U
1

0.05
0.1
2
T/I (m K/W)

0.15

Fig. 9.16: Characteristic curves of two thermal solar collectors with different thermal loss coefficients
A solar water heating system comprises, in addition to collectors, also a closed-loop piping system,
a heat exchanger, a tank to store the hot water, a pump as well as controls. Solar hot water heating is regarded today as a proven and economic technology.

Energy and Sustainability

172

Photovoltaics (PV) denotes the direct conversion of solar radiation into electric energy. The physical basis is the photovoltaic effect. The energy conversion takes place in the solar cells that are
bonded to the modules. The efficiency of todays commercially available solar modules is between
about 6 to 18 percent. If the PV system is connected to a grid, the direct current produced by the
solar cells is converted to an alternating current with a power inverter. PV systems are to date not
very prevalent in Swiss buildings mainly due to their relatively high costs.
When solar components are integrated in buildings, it can be attractive to directly build them into
the facade or roof. Ideally these elements can, in addition to producing heat or electricity, also take
on the function of weather protection. As a consequence other elements are not required anymore
and costs can be saved.

9.5

Total Energy Expenditure During a Life Cycle

The energy used for heating and possibly also for cooling of a building is more or less constant as
yearly consumption during the service life of a building. It can be designated as operating energy
or direct energy.
Additional energy is necessary particularly for the production of materials, but also for the construction of buildings, for repair and rehabilitation work as well as in the end for the building demolition
and waste disposal. This energy is referred to as indirect or embodied energy [9.13]. In contrast to
direct energy, the indirect energy doesnt accumulate continuously over the life cycle, but instead
only in specific phases.
For many materials the energy required for production is known from investigations of the manufacturing processes (Tab. 9.2).
Material

Embodied energy
(kWh/t)

aluminum
steel
plastics
glass
cement
mineral wool

34'183
8'138
14'650
5700
1'383
4'070

Tab. 9.2: Energy required to produce different materials [9.14]


In order to illustrate the significance of embodied energy the thermal insulation of a building is considered. The heat loss of a building in operation (direct energy) decreases with the increase in
thickness of the thermal insulation. The required energy for production (indirect energy) however
increases with the increasing thickness of the thermal insulation. What is then, energetically, the
optimal insulation thickness? The area-related energy loss through the thermal insulation during its
service life (direct energy) amounts to:

qdir = U n HDD 24 3600

J
2
m

service-life of thermal insulation (number of years)

thermal transmittance

W/m2K

HDD

heating degree days

Kd

The embodied energy contained in the thermal insulation (indirect energy) amounts to:

qind = x

J
2
m

Energy and Sustainability

173

specific embodied energy of the thermal insulation

J/m3

insulation thickness

The total energy q is the sum of the direct and indirect energy:
q = q dir + q ind

The total energy should be a minimum, thus:


dq
dx

=0
x = x opt

In very rough approximations it is assumed that the number of heating degree days is not dependent on the insulation thickness and the contributions of the heat transfer resistance in the thermal
transmittance U are negligible (U /x). The second assumption is very well fulfilled with larger insulation thicknesses. According to this, the optimal thickness of the thermal insulation is:

xopt =

n HDD 24 3600

(m )

The relationship between direct, indirect and total energy und insulation thickness are graphically
displayed in Fig. 9.17 (left). This figure also shows (right) that the energetically optimal insulation
thickness is larger the longer the service life is. For this the following numerical values are assumed:
location of Zurich:

HDD = 3260 Kd

mineral wool:

= 0.04 W/mK

= 80 kg/m3

and

1200
Optimal thickness of insulation (m)

direkt

Energy (MJ/m )

1000

indirekt
total

800
600
400
200
0
0

0.2
0.4
0.6
0.8
Thickness of insulation (m)

0.8
0.6
0.4
0.2
0

20

40
60
80
Service life (year)

100

Fig. 9.17: Direct, indirect and total energy as a function of the insulation thickness for n = 1 (left)
and optimal insulation thickness as a function of the service life of the thermal insulation (right).
The embodied (indirect) energy of the mineral wool is applied in accordance with Table 9.2. The
graphic shows that todays common insulation thicknesses are far below the energetically optimal
insulation thickness. This means that with todays common insulation thicknesses the total energy

Energy and Sustainability

174

needs are largely determined by the direct energy. Similar considerations can basically be carried
out for entire buildings.
Figure 9.18 shows the distribution of required indirect energy for the production and construction of
an office building [9.15]. The supporting structure contributes to nearly half of the total embodied
energy. The thermal insulation, the windows and doors as well as a portion of the building equipment are relevant for direct energy demand. These building components comprise up to about
30% of the total embodied energy.
interior finish 7%
HVAC 18%
windows, doors 14%
cladding 3%

construction site 9%

thermal insulation 1%

bearing structure 49%

Fig. 9.18: Embodied energy of a low-energy office building [9.15]


In Switzerland the average service life of buildings is about 80 years. The operating energy can be
summed up over the entire service life and compared with the embodied energy. Studies have
shown [9.15] that the portion of indirect energy in old buildings is only about 10% and about 20% in
newer. I.e. that the energy embodied in building elements relevant for operating energy is only responsible for about 3% to 6% of the total energy consumption! A greater investment of a few percent in embodied energy of building elements relevant for operating energy can however reduce
the operating energy requirements in certain circumstances by 50% and more. That is to say that
the investment in embodied energy for better thermal insulation, windows and building equipment
is in general practically negligible in comparison to the operating energy savings.
The consideration of embodied energy however becomes important with low- or zero-energy buildings. Some of todays zero-energy buildings (100% embodied energy) were made possible by the
application of technical equipment that contain considerable amounts of embodied energy and accordingly, environmental pollution. Therefore, an optimum as regards total energy, or still better, an
ecological optimum is something to strive for. Different studies indicate that this optimum lies relatively close to zero-energy for operation. In any case however, buildings should be constructed
with materials as little polluting as possible and using as little amounts of material as possible.

9.6

Energy Demand per Floor Area and Energy Standards

Figure 9.19 shows the energy flow diagram of a building with nomenclatures according to [9.9]. It
can be differentiated between the system boundaries for the heating demand (1), the heating demand for hot water (2), the heating and hot water system (3) as well as for the entire building (4).
The demands on energy consumption can either be formulated by end energy (EEl and Ehww) or by
the useful energy (Qh und Qww) (Tab. 9.1). In the Swiss standard SIA 380/1 for thermal energy in
buildings [9.9] the requirements are defined based on useful energy. The energy consumption per
year in MJ (= 106 Joule) is based on the so-called energy reference area (m2). The energy reference area is the sum of all floor areas both above and below ground that have to be heated or
cooled for the use. The requirements are different depending on the building category and geometry (Fig. 9.20). They are formulated either as limits (= mandatory to comply with) or target values (=

Energy and Sustainability

175

to strive for). Less strict requirements are given for refurbishments (= 140 % of the limit values).
The limit of heating demand as a function of the building geometry is according to the SIA 380/1
standard:

Qh,li = Qh,li 0 + Qh,li


Ath
AE

Ath
AE

(MJ/m2a)

thermally weighted building-envelope area


energy reference area

m2
m2

The thermally weighted building-envelope area Ath is defined as the outer surface area of the building in which areas adjacent to unheated rooms and adjacent to the ground are weighted less (for
details refer to [9.9]). That is, small values of Ath/AE correspond to large compact buildings and
large values of Ath/AE to small buildings (Fig. 9.20). The requirements are thus stricter for large
buildings because it is easier in these cases to achieve low consumption values.

end energy
1
2
3
4
EEl
Ehww
Qg
Qh
Qhww

useful energy

system boundary of heating demand


system boundary of heat demand for hot
water
system boundary of heating and hot water system
system boundary of building
energy demand for electricity
energy demand for heating and hot water
heat gains
heating demand
heat demand for heating and hot water

Qi
QiEl
QiP
QL
Qr
Qs
QT
Qtot
Qug
QV
Qww
WRG

internal heat gains


internal heat gains from electricity
internal heat gains from people
heat losses of heating and hot water
systems
gained environmental heat
solar heat gains
transmission heat losses
total heat losses
usable heat gains
ventilation heat losses
heat demand for hot water
heat recovery

Fig. 9.19: Energy flow diagram [9.9]

176

Energy and Sustainability

Building Category

Limits*
Qh,li0
Qh,li
2
MJ/m a MJ/m2a
80
90

residence MFH

II
III

residence SFH
office

90
85

90
95

IV
V

school
shop

100
75

95
95

VI
VII

restaurant
assembly hall

115
120

95
95

VIII
IX

hospital
industry

100
85

100
90

X
XI

warehouse
gymnasium

85
105

85
85

XII

indoor swimming pool

90

140

*at 8.5C annual mean exterior temperature

Fig. 9.20: Limits for the heating demand according to [9.9]; the target value is 60% of the limit value
The energy demand can of course also be expressed in in terms of end energy (Fig. 9.19). This
value is a quantity for the annual demand for energy per unit area for the provision of heat or electricity in a building (reference area = energy reference area).
Energy demand for heat (space heating + hot water): Ehww (MJ/m2a)
Energy demand for electricity: EEl (MJ/m2a)

Bemerkung: Werte gltig fr 20C Raumtemperatur

Tab. 9.3: Limit and target values for thermal transmittances U according to [9.9]
In the SIA 380/1 standard the requirements are either expressed in terms of heating demand per
floor area Qh (see above) or the thermal transmittance of building elements (Tab. 9.3). According

Energy and Sustainability

177

to [9.9] no heating demand calculation for the whole building (Sect. 9.2) must be made if limit values for thermal transmittances are fulfilled.
The energy demand per floor area of buildings increased significantly from the beginning of the
20th century up until the 1970s. Due to an increased push for savings, a reversal of the trend set in.
It is expected that this trend will continue in the next years (Fig. 9.21).

Fig. 9.21: SIA reduction path: Development of the energy demand per floor area for new buildings
For buildings with low heating demand, the required power for heating is also smaller (Tab. 9.4).
With low-energy buildings the (installed) power of internal sources can typically achieve the required heating demand (Tab. 9.5). That is to say, when the internal sources are in operation, the
demand for heating can thereby be covered. A possible heating system then has only a compensatory function.

Energy
2

Power
2

(MJ/m a)

(W/m )

old buildings

ca. 600

ca. 70

newer buildings

ca. 300

ca. 35

low-energy buildings

ca. < 100

ca. < 12

Tab. 9.4: Typical values for heating in terms of energy and power

lighting
people
appliances

Use

Installed Power
2
(W/m )

office
residence
office
residence
office
residence

10 15
46
8 10
26
15 50
2 10

Tab. 9.5: Typical values for internal heat sources [9.15]

Energy and Sustainability

178

Minergie and Minergie-P are quality labels for buildings that stand for high comfort with low energy
consumption and an economic construction [9.16]. Figure 9.22 gives an overview of the requirements on a building according to the Minergie and Minergie-P standards. With regards to construction costs, it is encouraged that any additional costs in construction, compared to conventional
construction, not be higher than 10% (Minergie), and 15% (Minergie-P).

Fig. 9.22: Comparison of Minergie and Minergie-P building properties [9.16]

9.7

Strategies for Low-Energy Buildings

Based on previous statements (see Chap. 8) basic strategies can be formulated for the design of
buildings with a very low energy demand for heating and cooling. In doing so a distinction can be
made between the building structure (= building envelope and interior building elements) and the
building equipment (= technical installations) as well as other factors.
Building Structure:
- continuous, good thermal insulation of transparent (U 0.9 W/m2K) and opaque (U 0.1
W/m2K) elements of the building envelope
- thermal-bridge-free building envelopes: load bearing structure on the warm side, no penetrations through the insulation layer with materials with a high thermal conductivity (metal, concrete, stone)
- air-tight building envelope (n50 0.6 1/h)
- solar gains in winter: large glazed areas preferably facing south, try to avoid shading from balconies, overhangs, etc.
- solar gains in summer reducible: best with external, variable shading devices (g < 0.15)

Energy and Sustainability

179

- active thermal-mass in building interior: e.g. concrete ceiling, brick walls etc. (elements not
thermally uncoupled by suspended ceilings, subfloors or carpets)
- building geometry optimized: small outer surface area (= heat loss area) in relation to building
volume; i.e., advantageous are large, compact buildings
Building Equipment (HVAC):
- mechanical ventilation with heat recovery (efficiency > 0.8)
- demand-regulated, optimized controls
- solar collectors (hot water, possibly also heating/cooling or photovoltaic)
- use geothermal heat (earth-to-air ground heat exchanger, heat pump)
- individual heating-cost billing
Other Factors:
- environmentally-friendly materials with little embodied energy
- correct user behavior: e.g. permanently opened (tilt-) windows can drastically increase the energy demand in a low-energy building, reasonable room temperatures etc.

9.8

Existing Building Stock and Refurbishment

Mainly due to the tightening of the SIA 380/1 standard as well as the regulations in the Cantons,
newly constructed buildings today require considerably less energy than existing. The analysis of
the building stock shows however that this only very slowly affects the total energy consumption
(Fig. 9.23). In Switzerland the average energy demand per floor area of buildings still amounts to
about 600 MJ/m2. The buildings that were constructed before circa 1980 primarily dominate the total consumption!

Fig. 9.23: Energy demand of residential buildings (data from [9.17])

180

Energy and Sustainability

Construction is a slow process. In Switzerland only about 1% to 2% of buildings are erected or refurbished annually. In contrast to this, for example, each year about 9% of the cars are replaced. If
a new energy-efficient technology comes on the market and is declared as mandatory, then todays automobile fleet can be replaced within approximately 20 years. By todays low building
modernization rate it takes however at least 100 years in order to replace todays building stock
(Fig. 9.24). If we decided today to construct, as of now, only zero-energy buildings, it would thus
take decades before the energy demand of the total building stock drops considerably. Thus in order to reduce the energy demand of buildings as quickly as possible not only strict energy regulations are important, but also the rate of building refurbishment must be boosted with suitable incentives.

Fig. 9.24: Rate of renewal and penetration curve [9.15]

9.9

Climate Change and Energy Demand

The gradual increase of temperatures due to climate change (cp. Sect. 2.5) means that the boundary conditions of buildings are slowly changing. Considered over a long period, this can lead to
significant changes in the energy balance of buildings. Taking the typical service life of buildings to
be about 80 years this also means that the buildings that are constructed today must also function
some day in a warmer climate.
Studies show [9.18] that in the period between 1901 to 2003 the heating degree days have decreased by 11% to 18%. During the same time period, the cooling degree days have increased by
50% to 170% (starting from low absolute values). This tendency will be even more pronounced in
the 21st century (Figures 9.25 and 9.26) so that it can be taken with high probability that in the future our buildings will have to be heated less but will need to be cooled more. In this study, temperatures measured by MeteoSchweiz [9.19] were employed for the past and temperature projections of the IPCC [9.20] were used for the future.

Energy and Sustainability

181

Heating degree days (Kd)

7000
6000
Davos
5000
4000

Zurich

Geneva

3000
2000
1000

Lugano
= 10C
g

0
1900

1950

2000
Year

2050

Fig. 9.25: Measured (1901 2003) and projected development of heating degree days [7.18]
1200
Cooling degree days (Kd)

1000
800

bal

= 18.3C

Lugano
Geneva

600

Davos

Zurich

400
200
0
1900

1950

2000
Year

2050

Fig. 9.26: Measured (1901 2003) and projected development of cooling degree days [7.18]

9.10 Summary
"Sustainable development meets the needs of the present without compromising the ability of future generations to meet their own needs" (Brundtland-Commission, 1987). Sustainable construction means to minimize the flow of energy and materials during the entire life cycle of a building
(construction, operation, refurbishment, demolition). A high priority has to be given to the reduction
of the energy demand during operation. In comparison to typical old buildings, the heating demand
in buildings can today be reduced by a factor of 5 to 10 with an economically justifiable investment.
Higher costs can be outweighed by better comfort surface temperatures closer to the interior air
temperature as well as better air quality due to controlled ventilation. To strive for is to at least partially cover the remaining energy demand with renewable energy.
The basic principles and strategies for buildings with very low energy demand have been presented. The practice shows that very large architectonic diversity is also possible with low-energy buildings.

Energy and Sustainability

182

Problems
Problem 1: Energy Flows in a Room
Given is an interior room of a building (drawing). The floor, ceiling and interior walls are adiabatic,
i.e. heat can flow only through the exterior wall. A steady state can be assumed.

I,

U2, A2, F, g

PAP
Ph

U1, A1

a.) Provide a balance of all energy flows.


b.) How much heating power is required so that i = 20C = constant? Use the following:
exterior climate:
I = 200 W/m2, e = -5C
exterior wall:
U1 = 0.3 W/m2K, A1 = 7 m2,
window:
U2 = 2 W/m2K, A2 = 3 m2, Fg = 0.4
heat output from people and appliances:
PAP = 150 W
air-flow rate, interior exterior:
V = 30 m3/h
Problem 2: Transmission and Ventilation Heat Losses of an Apartment
An apartment in a multi-family house in Zurich has a volume of V = 250 m3 and a portion of the facade area of A = 120 m2. The air exchange rate is n = 0.5/h. Heat flows to the adjacent rooms in
the building interior can be neglected.
a.) What can the maximum average thermal transmittance of the facade U be, in order that the
transmission heat losses will not be greater than the ventilation heat losses?
b.) Relatively, how much energy could be saved from transmission and ventilation losses, if the
apartment was heated to an interior temperature of only 18C during the heating period instead of 20C? This assessment should be conducted for a heating limit of g = 12C (Zurichcity: HGT20/12 = 3260 Kd/a and HT12 = 208 d/a).
Problem 3: Transmission and Ventilation Heat Losses SFH vs. MFH
Given are two cubic shaped buildings, a single family house (SFH) (edge length s1 = 9 m) and a
multi-family house (MFH) (edge length s2 = 30 m), that both exhibit an identical average thermal
transmittance U = 0.3 W/m2K for the building envelope (incl. floor) and the same air exchange rate
of n = 0.5 h-1 (Assume: ground temperature = exterior air temperature).
a.) What is the ratio between transmission and ventilation heat losses for both buildings?
b.) Describe your results in a short sentence.

Literature
[9.1]

Meadows D. L., Meadows D. H., Zahn E., Milling P., The Limits to Growth, Universe Books,
New York, 1972

Energy and Sustainability

183

[9.2]

Meadows D. H., Meadows D. L., Randers J., Beyond the Limits, Chelsea Green Publishing
Company, Vermont, 1992

[9.3]

M. Wackernagel, zitiert in Limits to Growth: The 30-Year Update, D. Meadows et al.,


Chelsea Green Publishing Company, 2004

[9.4]

Schweizerische Gesamtenergiestatistik 2005, Bundesamt fr Energie, Bern, 2006

[9.5]

World Energy Outlook 2006, International Energy Agency, Paris, 2006

[9.6]

Analyse des schweizerischen Energieverbrauchs 2000-2006 nach Verwendungszwecken,


Bundesamt fr Energie, Bern, 2008
[9.7] Glicksman L.R., Energy efficiency in the built environment, Physics Today, July Issue, 2008
[9.8] Impulsprogramm Bau - Erhaltung und Erneuerung, Eidgenssische Drucksachen- und Materialzentrale, Bern, 1991
[9.9] SIA 380/1, Thermische Energie im Hochbau, Hrsg: Schweizerischer Ingenieur- und Architekten-Verein, 2009
[9.10] EN 832, Wrmetechnisches Verhalten von Gebuden Berechnung des Heizenergiebedarfes Wohngebude, CEN, Brssel, 1998
[9.11] ASHRAE Fundamentals Handbook 2001 (SI Edition), Chapter 31, American Society of
Heating, Refrigerating, and Air-Conditioning Engineers, Atlanta, GA, 2001
[9.12] Adnot J. et al., Energy Efficiency and Certification of Central Air Conditioners, Study for
D.G. Transportation-Energy (DGTREN) of the Commission of the EU, Final report, Armines,
Paris, 2003
[9.13] Spreng D., Graue Energie, vdf Hochschulverlag, Zrich, 1995
[9.14] Kohler N., Analyse nergtique de la construction, EPF-Lausanne, 1986
[9.15] Keller B., Bauphysik: Die Energetik des Gebudes, Vorlesungsskript ETH, Zrich, 2006
[9.16] Verein Minergie, Bern; http://www.minergie.ch, 2010
[9.17] Energieplanungsbericht 2002 fr den Kanton Zrich, Bericht des Regierungsrates ber die
Energieplanung, AWEL, Abteilung Energie, Zrich, 2003
[9.18] Christenson M., Manz H., Gyalistras D., Climate warming impact on degree-days and building energy demand in Switzerland, Energy Conversion and Management, Vol. 47, 2006,
671-686
[9.19] Begert M., Seiz G., Schlegel T., Musa M., Baudraz G., Moesch M., Homogenization of
measured time series of climatic parameters in Switzerland and computation of norm values 1961-1990. Final project report NORM90, MeteoSwiss, Zurich, 2003
[9.20] Intergovernmental Panel on Climate Change: http://www.ipcc.ch. Retrieved Octo

Chapter 10
Daylight

Daylight

186

10

Daylight

In prehistoric times humans lived out in the open air and resided in their caves primarily at night.
The light that penetrated through the cave opening during the day highlighted the cave exit, but only produced a very poor light within. Today many people in the so-called industrialized countries
spend about 90% of their time indoors.
The good daylighting of interior spaces brings multiple benefits:
- psychological & physiological: It creates a connection to the outside world. The weather, or at
least the outside light situation, can also be perceived in the interior.
- ecological & economical:
- If daylight is used, then artificial lighting and with that electrical energy can be saved.
- The use of artificial lighting in summer can contribute to an overheating of the building. In the
case of excessive heat the thermal comfort in the interior can therefore be improved with
daylighting (better light output, refer to Section 10.2). In the case of an air-conditioned building, through the use of daylight, in addition to the energy savings with lighting, the energy to
operate the refrigerator unit can be reduced. In winter the released heat from the electrical
lighting is however of use: the artificial light can also be considered as expensive supplementary heating.
- Architectural: "There is no architecture without light", once remarked the physicist J.M.
Waldram. Daylight is a design element of architecture.

10.1 Solar Radiation and Spectral Luminous Efficiency of the Human Eye
During the course of evolution the human eye developed its sensitivity in the same wavelength
range in which the solar radiation has the highest intensity (Fig. 10.1).

Fig. 10.1: Spectral distribution of solar radiation [10.1] and spectral sensitivity of the human eye
(photopic vision).
Visible light is understood as electromagnetic radiation in the wavelength interval of 380 to 780 nm,
in which the human eye is sensitive. For high intensities (photopic vision V(); normal, daytime) the
maximum sensitivity of the eyes lies at a wavelength of 555 nm (yellow-green). With lower intensi-

Daylight

187

ties (scotopic vision V(); low illuminance levels, nighttime) the maximum sensitivity shifts to the
wavelength 507 nm (Fig. 10.2).

Fig. 10.2: Spectral sensitivity of the human eye (photopic and scotopic vision)[10.2].

10.2 Fundamental Photometric Terms and Relationships


10.2.1 Luminous Flux
The amount of visible light emitted by a source is referred to as luminous flux with the unit lumen (lm). The luminous flux corresponds to a certain energy flow, or power, in Watt.
The conversion of power into luminous flux can be carried out by means of the spectral radiative
flux e and the spectral luminous efficiency V() (Fig. 10.3). This is referred to as photometric
weighting.

= Km

780 nm

el

( l ) V ( l ) dl

( lm )

K m = 683 lm / W

380 nm

The scaling factor Km is called maximum luminous efficacy.


The same power produces e.g. with 450 nm (blue-violet) much less lumens than with 555 nm (yellow-green), the maximum spectral sensitivity. Monochromatic light of a 555 nm wavelength corresponds to the maximum luminous efficacy of 683 lm/W. Daylight has a luminous efficacy of about
90 to 100 lm/W. The luminous efficacy of daylight is not constant, but depends on the cloud condition, the solar elevation angle and on whether the solar radiation is direct or diffuse (Fig. 10.4). The
lower the solar elevation angle is, the redder the light is and so the lower is the luminous efficacy.
Table 10.1 shows the luminous efficacy for different light sources. The luminous efficacy in lumen
per watt can also understood as visual performance of a light source.

188

Daylight

Fig. 10.3: Determination of the spectral luminous flux for a given spectral radiative flux e by
means of the spectral luminous efficiency of the human eye V()

Fig. 10.4: Luminous efficacy as a function of the solar elevation angle [10.3]

Daylight

189

Light Source

Visual Performance

daylight

ca. 90 100 lm/W

incandescent light-bulb

ca. 14 lm/W

halogen lamp

ca. 25 lm/W

fluorescent tube

ca. 40 90 lm/W

light-emitting diode (LED)

ca. 30 200 lm/W

sodium-vapor lamp

ca. 90 145 lm/W

monochromatic light ( = 555 nm)

683 lm/W

Tab. 10.1: Visual performance of light sources


10.2.2 Illuminance E
The illuminance E is a measure of the luminous flux incident per surface area A and is given in
the unit lux (lx).
E=

lm
= lx )
m2

In the case of light striking a surface at an oblique angle, the angle of incidence must then be considered according to cosine law.
The illuminance on a horizontal surface varies depending on the solar elevation angle and the atmospheric conditions (Fig. 10.5). Table 10.2 gives the illuminance outdoors for different conditions.
For daylight it can approximately be taken that: 1 W/m2 100 Lux. Table 10.3 gives the illuminance levels required to perform different visual tasks.

Setting

Illuminance E (lx)

direct irradiation

60'000 100'000

cloudy summer day

20'000

high fog cover in winter

10'000

cloudy winter day

3'000

full moon night

0.25

new moon, starlight

0.01

Tab. 10.2: Illuminance outdoors [10.3]

Visual Task

Illuminance E (lx)

orientation, only temporary presence (e.g. circulation area)

50

easy visual task (e.g. restaurant)

200

normal visual task (e.g. home, office)

500

difficult visual task (e.g. office, laboratory, workshop, retail store)

1000

very difficult visual task (e.g. precision assembly, watchmaking)

2000

Tab. 10.3: Illuminance levels required to perform different visual tasks

Daylight

190

Fig. 10.5: Course of horizontal illuminance outdoors in Zrich [10.2]


10.2.3 Luminous Intensity I
Many light sources emit the luminous flux in different intensities in different directions (directiondependent radiation characteristic). The luminous intensity I specifies which luminous flux falls in
a certain solid angle and is given in the unit candela (cd).
I=

lm
= cd )
sr

Fig. 10.6: Solid angle

Daylight

191

The solid angle is defined by the ratio of the area of a spherical calotte to the square of the sphere
radius (Fig. 10.6):

A
r2

( sr )

The unit of the solid angle is called steradian (sr). The total surface area of a sphere corresponds
to a solid angle = 4. The correlation between luminous flux , luminous intensity I and illuminance E is shown in Figure 10.7. The luminous intensity corresponds to the luminous flux that falls
in a certain solid angle. For example, with a headlamp the luminous flux is focused. That is, the
light is bounded by a small solid angle. So, with a given luminous flux, a much higher luminous intensity is generated. The illuminance E describes the effect on the target, i.e. how intense the light
acts on a surface.

Fig. 10.7: Relationship between luminous flux , luminous intensity I and illuminance E
10.2.4 Photometric Inverse-Square Law
A light source with a luminous intensity I lights up an area A at a distance r from the light source.
The linear extent of the surface A is small in comparison to the distance r. The surface A determines the following solid angle:

A
r2

The luminous flux that strikes the surface A is:

= I = I

A
r2

The illuminance E on the surface A amounts to:


E=

I
r2

The illuminance E that is produced by a given light source with the luminous intensity I thus increases inversely proportional to the square of the distance r (photometric inverse-square law).

Daylight

192

10.2.5 Luminance L
A light source displays a surface area. If one refers to the luminous intensity of the source on their
surface A (projection perpendicular to the visual line), one obtains the luminance of the source:
L=

I
A

cd
2
m

Luminance is the quantity that is relevant for human sight. Whether an object can be recognized,
depends on the luminance contrast to the surroundings. The physiological contrast is defined as:

Studies have shown that the human eye exhibits a maximum contrast sensitivity, and accordingly
visual performance, with a background luminance of about 200 to 2000 cd/m10.
Each illuminated surface turns itself into a light source by its reflectance. The illuminance E that
falls on an area together with its reflectivity determines the luminance, under which the area appears to the eye. The luminance of a surface in the case of an ideal diffuse reflection is:
L=

Given that the human eye needs a luminance at a minimum of 200 cd/m2 for good visual performance, the required illuminance, for white paper ( = 0.8) amounts to (cp. Table 10.3):
E=

L 200
= 785 lx
=
0 .8

Fig. 10.8: Luminous flux , illuminance E and luminance L


10.2.6 Overview
In summary one can say:
- The luminous flux describes the total output of the light source.
- The illuminance E describes how much the incident light illuminates the lit object, that itself becomes a source of light with a luminance L as a result of reflection.

Daylight

193

- Luminance L and luminous intensity I describe the light source (lamp as well as the illuminated
object).

10.3 Luminance Distribution of the Sky


To achieve a better comparability, one uses with daylight planning a so-called standard sky. The
standard sky corresponds to a uniformly overcast sky, i.e. one assumes that 100 percent of the sky
is covered with clouds. This sky condition is, precisely in mid-winter, unfortunately a relatively
common situation on the Swiss plateau. It is also the sky condition with which the use of daylight is
critical. If a lot of direct solar radiation exists then often rather too much daylight exists and glare
effects can occur.
The CIE-standard sky is defined by the Commission Internationale de lEclairage, or International
Commission on Illumination, and is given as the relative luminance distribution of the illuminant
sky (Fig. 10.9). denotes the sun elevation angle.
L( ) =

1 + 2 sin
L90
3

According to this distribution the luminance is three times larger at the zenith ( = 90) than at the
horizon ( = 0).
LZenit
=3
LHorizont

Fig. 10.9: Relative luminance distribution of the CIE-standard sky


The result of this luminance distribution is that the illuminance on a vertical surface amounts to just
barely 40% of the illuminance on a horizontal surface.
E vertikal
= 0.397
E horizontal

Therefore, skylights utilize daylight significantly more efficient than vertical glazings.

Daylight

194

10.4 Transmittance Factors of the Building Envelope


How much daylight can reach into an interior room depends on the transmission properties of the
building envelope. The visible light transmittance of a glazing can be calculated by a weighing of
the spectral transmittance properties of the glazing with the spectral luminous efficiency of the human eye [10.4]:
780 nm

V =

( ) V ( ) d
()

380 nm
780 nm

D V ( ) d

380 nm

() denotes the wavelength-dependent transmittance of the glazing, V() spectral luminous efficiency of the eye and D the relative spectral distribution of standardized light [10.4].
Table 10.4 gives typical visible light transmittances of glazings. To provide interiors well with daylight, high values of V are desirable.

visible light transmittances v (-)


single glazing

0.9

double glazing

0.82

insulating glazing unit (double)

0.77

insulating glazing unit (triple)

0.65

solar control glazing

0.2 0.6

Tab. 10.4: Typical visible light transmittances of glazings


With regards to the different demands daylight, glare protection, solar control in summer and
passive use of solar energy in winter often arise conflicting requirements. Today the different
demands are accommodated mostly with variable additional elements external shading and internal glare protection devices. Electrochromic devices have to date still not achieved an appreciable prevalence.
A substantial portion of window surfaces is often taken up by frames and muntins that reduce the
light transmission. This can be accounted for with a reduction factor K1 which considers the area
share in the wall aperture.

K1 = 1

AFrame
AWall Aperture

( )

Depending on air pollution countryside or densely populated industrial areas as well as cleaning schedules light transmission is further reduced. The reduction factor K2 accounts for dirt (K2 =
ca. 0.5 0.9).
Depending on the frame depth and height, an angle-dependent frame shading can occur as well
(Fig. 10.10) that can be accounted for with a reduction factor K3.

Daylight

195

Fig. 10.10: Reduction factor K3 due to frame shading

10.5 Daylight Factor


In order to be able to evaluate the illumination of a room by daylight, the illuminance on a horizontal surface in the room interior E1 at a defined point (e.g. table surface) is compared to the illuminance on a horizontal surface outdoors E0 for an overcast sky (CIE-standard sky). The ratio that is
composed of these two illuminances is called daylight factor TLQ (Fig. 10.11).
TLQ =

E1
100
E0

(%)

Fig. 10.11: Definition of the daylight factor TLQ


10.5.1 Components of Daylight Factor
The daylight factor can be described as follows (Fig. 10.12):
TLQ = (TH + TV + TR ) V K 1 K 2 K 3

(%)

TH

sky component (direct from sky)

TV

externally reflected component (reflection of opposite lying obstructions and/or ground)

TR

internally reflected component (reflection of internal surfaces)

Daylight

196

light transmittance of glazing

K1

reduction factor for frame and muntins

K2

reduction factor due to dirt

K3

reduction factor due to frame shading (non-perpendicular light incidence)

Fig. 10.12: Components of daylight factor


The larger the daylight factor is, the brighter a room appears and the better connected one feels
with the outdoor environment (Fig. 10.13).

Fig. 10.13: Perception of daylight factors [10.3]


The sky component TH arises from the visible patch of sky through the window. The qualitative
course of the sky component of the daylight factor as a function of the distance from the window in
the interior is evident in Figure 10.14. The light colored portion of the hemisphere on the horizontal
surface corresponds to the direct sunlight from a patch of sky visible at the point considered, and
becomes smaller with increasing distance from the window.

Daylight

197

Fig. 10.14: Dependency of the sky component of the daylight factor on the interior location in the
absence of obstructions
In simple cases, as for example with a vertical glazing, the daylight factor can be manually (roughly) calculated. This method will be introduced in the following. In complicated cases daylight factors
can be determined with the aid of computer programs.
10.5.2 Sky Component TH
The sky component TH can be determined quantitatively with the aid of the so-called skylight or
Waldram diagram. This diagram (Fig. 10.15) depicts a projection of half of the celestial dome (corresponding to TH = 50%). The division was so chosen that every little square in the grid constitutes
the same illuminance on a horizontal surface. The procedure is as follows:
- From the reference point in the room one extends the perpendicular line to the plane of the window and determines the angles to the two side edges as well as upper and lower edges of the
window (azimuth angle left Fl and right Fr , elevation angle low F1 and high F2).
- One plots the angle-picture in the Waldram diagram, utilizing the mesh lines for the horizontal
and vertical window outline.
- One determines analogously possible obstructions with respect to their silhouette within the
lines of the window angle-picture and plots it likewise in the diagram.
- One counts out the contained squares from the angle-picture of the window minus the angle
picture of the obstruction (visible sky) (nH).
For the sky component TH it amounts to:
TH = nH 0.05

(%)

10.5.3 Externally Reflected Component TV


One determines analogously as with TH the contained number of houses nV from the angle-picture
of the obstruction. The contribution to the daylight factor is now co-determined by the reflective
power of the obstruction (Table. 10.5). The luminance of the portion that corresponds to the obstructed area, is less than the one of the sky and the contribution to the daylight factor is reduced.
TV = nV 0.05

(%)

Often an average value of = 0.15 is assumed for the reflectance.

198

Daylight

Fig. 10.15: Projection of half of the celestial dome: Waldram diagram

Tab. 10.5: Reflectances of building materials and surfaces [10.3]


10.5.4 Internally Reflected Component TR
The internally reflected component TR can be approximately determined with the Hopkinson formula:

Daylight

TR =

AF
1

(fo BW + fu DW ) 100
AR (1 m )

199

(%)

m2

AF

window area (wall aperture)

AR

total room surface areas (floor, ceiling, walls)

average reflectance of all room surfaces

m2
-

reflectance of windows: F = 0.1

BW

DW

average reflectance of floor and lower part of walls


(lower half from mid-height of window) without window wall

average reflectance of ceiling and upper part of wall


(upper half from mid-height of window), without window wall

fo, fu

window factors, according to Fig. 10.16

obstruction distance angle, measured from mid-height of window

200

Daylight

Fig. 10.16: Window factors fo, fu as a function of the obstruction distance angle
10.5.5 Example to Calculating the Daylight Factor
Figure 10.17 shows a room with a floor area 4 m x 5 m and a window size of 1.4 m x 3.6 m. For
this layout the daylight factor at the level of the windowsill is calculated at 1 m into the room, and
respectively at 4 m. The Figures 10.18 to 10.20 illustrate the approach to the solution.

Daylight

Fig. 10.17: Layout to the example [10.3]

201

202

Daylight

Fig. 10.18: Section and plan to the example [10.3]

Daylight

203

Fig. 10.19: Waldram diagram to the example for 1 m into the room (above) and 4 m (below) [10.3]

204

Daylight

Fig. 10.20: Calculation of the daylight factor (example) [10.3]

Daylight

205

10.6 Influence of Fenestration


The position of a window, not only its size, influences considerably the resulting daylighting conditions in the interior. In particular the light that falls nearly vertical can produce high illuminance on
the horizontal surfaces, because not only is the luminance highest at the zenith (CIE standard sky),
but also the cosine law has an effect (Fig. 10.21). Therefore, skylights are more effective than vertical windows.

Fig. 10.21: Three openings of different sizes that produce the same illuminance on a horizontal
surface.

Fig. 10.22: Influence of the vertical window position on the course of the daylight factor with singlesided illumination [10.2].

Fig. 10.23: Double-sided illumination of a room [10.2]

206

Daylight

The influence of the vertical position of the window on the course of the daylight factor in the case
of a single-sided illumination of a room is shown in Figure 10.210. With higher positioned windows
a more uniform illumination can be achieved. The usable room depth increases. A further improvement can be gained by a double-sided illumination (Fig. 10.23). The illuminances overlap and
the minimal value shifts in the direction of the center of the room.
Figure 10.24 shows how rooms can be very efficiently and uniformly illuminated with skylights.

Fig. 10.24: Impact of the type of skylight opening on the course of the daylight factor ( = window
area/floor area = 1/6). In addition, for TLQmin = 5% the required -value is indicated [10.5]
Fig. 10.25 shows the influence of an overhang on a vertical and horizontal surface, respectively.

Daylight

207

Fig. 10.25: Influence of an overhang on the vertical and horizontal surfaces, respectively [10.6]

10.7 Rules for Good Daylighting


In the following, a few rules for daylighting shall be formulated.
Skylights are especially efficient with respect to daylighting. Solar control and glare protection can
however also be required here.
With vertical glazings applies:
- The clear room height determines how much light, at the most, can enter the room: the greater
the room height, the more light.
- The proportion of the room height to the room depth determines the supply of daylight: A large
room height and a small room depth is beneficial.
- The maximum room depth that can be illuminated is ca. 5 6 m with conventional room heights.
- A large area glazing with a high transmittance is beneficial.
- The lintel should be as small as possible. That is, high-lying windows are favorable.

208

Daylight

- The glazing region below the work level provides only a small contribution to the daylighting, but
can have considerable consequences for the cooling load.
- Inner surfaces should be preferable light colored.
- A second sidelight can improve the daylighting, in particular the illuminances becomes more
uniform.
- Disadvantageous are fixed shading devices or balconies over the window. They especially
shield the direct skylight and reduce the daylight factor, particularly near the window.

10.8 Daylight Planning


Rules of thumb and Waldram diagrams arent always adequate for daylight planning. In order to
allow for reliable information about the daylighting conditions in rooms already in the planning
phase, computer programs are made available, with which on the basis of calculations with pictorial display of the results can do the same work with much more detail. A strength of powerful
programs is the photo-realistic rendering of the results. Thanks to such visuals also laypersons can
quickly imagine the daylight situation in a room (Fig. 10.26). In such representations also lines of
equal illuminance or equal daylight factors can be incorporated.

Fig. 10.26: Daylight simulation of an atrium


Daylight experiments with models of the relevant building parts can also be carried out. The illumination can thereby take place through natural or artificial sunlight (Fig. 10.27). With the help of
sensors in the model and photo-cameras quantitative as well as qualitative analysis can be made.
By means of Fig. 10.28 the daylight contribution with a given daylight factor and certain required
illuminances at a workplace (winter: 7.00 17.00 MEZ; summer 8.00 18.00 MESZ) can be estimated. For example artificial light can be turned off for a required illuminance of 500 lx with a daylight factor of 5% during 50% of the time.

Daylight

209

Fig. 10.27: Laboratory set-up to the investigation of daylighting conditions in buildings with direct
(left) and diffuse radiation (right)) [10.7]

Fig. 10.28: Times of operation solely with daylight for different daylight factors and required illuminances [10.6]

Daylight

210

Problems

Problem 1: Determination of the Daylight Factor


The daylight situation at the specified point P in a room is defined by the information given below:

Window

Reflectivity

azimuth angle: Fr = 35, Fl = 35


elevation angle: F = 40

walls: W = 0.8

Obstruction
azimuth angle: Vr1 = 15, Vr2 = 20
elevation angle: V = 35
obstruction distance angle: = 15

ceiling: D = 0.8
floor: B = 0.2
window: F = 0.1
reflectance of obstruction: V = 0.2

visible transmittance of glazing: V = 0.75


reduction factor for frame and muntins: K1 = 0.7
reduction factor for dirt: K2 = 0.8
A perspective of the interior, a section and a plan with the location of P can be found in the following pages. The drawings are not to scale! Using the Waldram diagram determine:
a)

the sky component TH at point P

b)

the externally reflected component TV at point P

c)

the internally reflected component TR at point P

d)

the daylight factor TLQ at point P

e)

the percentage share of the sky component, externally reflected component and internally reflected component to the TLQ (=100%), that has been calculated in d).

Problem 2: Sheens on Computer Screens


Annoying sheens on computer screens are caused by differences in luminance. This results in the
lights or irradiation mirroring themselves on the monitor screen. Think about measures to prevent
the annoying sheen and nevertheless maintain an acceptable illuminance:
a)

for providing daylight

b)

for artificial light

Problem 3: Photometric Fundamental Terms


A point light source has a power of 10 W and a luminous efficacy of 60 lm/W. The light is emitted
with a solid angle of 1 sr. Calculate the maximum illuminance on two surfaces at a distance of 1 m
and 3 m.

Daylight

Geometry to problem 1:

211

212

Waldram diagram (to problem 1):

Daylight

Daylight

213

Problem 4: Daylight factor in the office


skylight

window

Sketch the course of the average daylight factor


TLQ qualitatively at desk-height as a function of
the distance from window x for the following cases:
a.) Only the skylight exists.
b.) Only the window exists.
c.) Both openings exist.

Literature
[10.1] Iqbal M., An Introduction to Solar Radiation, Academic Press, Toronto, 1983
[10.2] Zrcher C., Frank T., Bauphysik: Bau und Energie, vdf, Zrich, 2004
[10.3] Keller B., Bauphysik: Stdtebaulich relevante Faktoren, Vorlesungsskript ETH, Zrich, 2006
[10.4] EN 410, Glas im Bauwesen Bestimmung der lichttechnischen und strahlungsphysikalischen Kenngrssen von Verglasungen, Europisches Komitee fr Normung, Brssel, 1998
[10.5] Fischer U., Tageslichttechnik, Rudolf Mller, Kln, 1982
[10.6] Keller B., Rutz S., Pinpoint Fakten der Bauphysik zu nachhaltigem Bauen, vdf Hochschulverlag, Zrich, 2007
[10.7] Scartezzini J.-L., ETH Lausanne, LESO Solar Energy and Building Physics Laboratory, ,
http://leso.epfl.ch/e/research_dl_scansky.html, 2010
[10.8] Haas-Arndt D., Ranft F., Tageslichttechnik in Gebuden, C.F. Mller Verlag, Heidelberg,
2007

Appendix

216

Appendix

A Weather Data
Tabelle 1: Globalstrahlung und Monatsmitteltemperatur einiger Standorte (Quelle: SIA 381/2, 1991)

Appendix

Tabelle 1 (Fortsetzung)

217

218

Tabelle 1 (Fortsetzung)

Appendix

Appendix

Tabelle 1 (Fortsetzung)

219

220

Tabelle 1 (Fortsetzung)

Appendix

Appendix

Tabelle 1 (Fortsetzung)

221

222

Tabelle 1 (Fortsetzung)

Appendix

Appendix

B Properties of Building Materials


Tabelle 2: Rechenwerte von Dmmstoffen fr bauphysikalische Nachweise (Quelle: SIA 279, 2004)

223

Appendix

224

Tabelle 3: Wrmeschutztechnische Bemessungswerte fr Baustoffe, die gewhnlich bei Gebuden


zur Anwendung kommen (Quelle: SN EN 12524:2000)
Stoffgruppe oder Anwendung

Rohdichte

kg/m
Asphalt

Bemessungs- Spezifische
wrmeleit- Wrmespeich
fhigkeit
erkapazitt

cp
W/(mK)
J/(kgK)

Wasserdampfdiffsionswiderstandszahl

trocken

feucht

2100

0,70

1000

50000

50000

als Stoff
Membran / Bahn

1050
1100

0,17
0,23

1000
1000

50000
50000

50000
50000

Mittlere Rohdichte

1800
2000
2200
2400
2300
2400

1,15
1,35
1,65
2,00
2,3
2,5

1000
1000
1000
1000
1000
1000

100
100
120
130
130
130

60
60
70
80
80
80

1200
1700
270

0,17
0,25
0,10

1400
1400
1400

10000
10000
10000

10000
10000
10000

120
200
<200
>400
200
1200

0,05
0,06
0,05
0,065
0,06
0,17

1300
1300
1500
1500
1300
1400

20
20
20
40
5
1000

15
15
10
20
5
800

1,23
1,95
1,70
6,36
3,56
5,68

0,025
0,014
0,017
0,013
0,0090
0,0054

1008
820
519
614
245
160

1
1
1
1
1
1

1
1
1
1
1
1

2500
2200
2000

1,00
1,40
1,20

750
750
750

920
900
100
200
300

2,30
2,20
0,05
0,12
0,23

2000
2000
2000
2000
2000

500
1000
990
970

0,60
0,60
0,63
0,67

2000
4190
4190
4190

2800
8700
8400
8900
7500
11300
7800
7900
7200

160
65
120
380
50
35
50
17
110

880
380
380
380
450
130
450
460
380

Bitumen

Beton (a)

Hohe Rohdichte
Armiert (mit 1% Stahl)
Armiert (mit 2% Stahl)
Fussbodenbelge
Gummi
Kunststoff
Unterlagen, porser Gummi oder
Kunststoff
Filzunterlage
Wollunterlage
Korkunterlage
Korkfliesen
Teppich / Teppichbden
Linoleum
Gase
Luft
Kohlendioxid
Argon
Schwefelhexafluorid
Krypton
Xenon
Glas
Natronglas (einschliesslich Floatglas)
Quarzglas
Glasmosaik
Wasser
Eis bei 10 C
Eis bei 0 C
Schnee, frisch gefallen (< 30 mm)
Neuschnee, weich (3070 mm)
Schnee leicht verharscht
(70100 mm)
Schnee, verharscht (< 200 mm)
Wasser bei 10 C
Wasser bei 40 C
Wasser bei 80 C
Metalle
Aluminiumlegierungen
Bronze
Messing
Kupfer
Gusseisen
Blei
Stahl
Nichtrostender Stahl
Zink
(a)

Die Rohdichte von Beton ist als Trockenrohdichte angegeben.

Appendix

225

Tabelle 3 (Fortsetzung)
Stoffgruppe oder Anwendung

Rohdichte

kg/m
Massive Kunststoffe
Akrylkunststoffe
Polykarbonate
Polytetrafluorethylenkunststoffe (PTFE)
Polyvinylchlorid (PVC)
Polymethylmethakrylat (PMMA)
Polyazetatkunststoffe
Polyamid (Nylon )
Polyamid 6.6 mit 25% Glasfasern
Polyethylen /hoher Rohdichte
Polyethylen/niedriger Rohdichte
Polystyrol
Polypropylen
Polypropylen mit 25% Glasfasern
Polyurethan (PU)
Epoxyharz
Phenolharz
Polyesterharz
Gummi
Naturkautschuk
Neopren (Polychloropren)
Butylkautschuk, (Isobuten-Kautschuk),
hart/heiss geschmolzen
Schaumgummi
Hartgummi (Ebonit), hart
Ethylen-Propylenedien, Monomer
(EPDM )
Polyisobutylenkautschuk
Polysulfid
Butadien
Dichtungsstoffe, Dichtungen und
wrmetechnische Trennungen
Silicagel (Trockenmittel)
Silikon ohne Fllstoff
Silikon mit Fllstoffen
Silikonschaum
Urethan-/Polyurethanschaum (als
wrmetechnische Trennung)
Weichpolyvinylchlorid (PVC-P) mit
40% Weichmacher
Elastomerschaum, flexibel
Polyurethanschaum (PU)
Polyethylenschaum
Gips
Gips
Gips
Gips
Gips
(b)
Gipskartonplatten
Putze und Mrtel
Gipsdmmputz
Gipsputz
Gipsputz
Gips, Sand
Kalk, Sand
Zement, Sand
(b)

Bemessungs- Spezifische
wrmeleit- Wrmespeich
fhigkeit
erkapazitt

cp
W/(mK)
J/(kgK)

Wasserdampfdiffsionswiderstandszahl

trocken

feucht

1050
1200
2200
1390
1180
1410
1150
1450
980
920
1050
910
1200
1200
1200
1300
1400

0,20
0,20
0,25
0,17
0,18
0,30
0,25
0,30
0,50
0,33
0,16
0,22
0,25
0,25
0,20
0,30
0,19

1500
1200
1000
900
1500
1400
1600
1600
1800
2200
1300
1800
1800
1800
1400
1700
1200

10000
5000
10000
50000
50000
100000
50000
50000
100000
100000
100000
10000
10000
6000
10000
100000
10000

10000
5000
10000
50000
50000
100000
50000
50000
100000
100000
100000
10000
10000
6000
10000
100000
10000

910
1240
1200

0,13
0,23
0,24

1100
2140
1400

10000
10000
200000

10000
10000
200000

60 - 80
1200
1150

0,06
0,17
0,25

1500
1400
1000

7000

6000

7000

6000

930
1700
980

0,20
0,40
0,25

1100
1000
1000

10000
10000
100000

10000
10000
100000

720
1200
1450
750
1300

0,13
0,35
0,50
0,12
0,21

1000
1000
1000
1000
1800

5000
5000
10000
60

5000
5000
10000
60

1200

0,14

1000

100000

100000

60 - 80
70
70

0,05
0,05
0,05

1500
1500
2300

10000
60
100

10000
60
100

600
900
1200
1500
900

0,18
0,30
0,43
0,56
0,25

1000
1000
1000
1000
1000

10
10
10
10
10

4
4
4
4
4

600
1000
1300
1600
1600
1800

0,18
0,40
0,57
0,80
0,80
1,00

1000
1000
1000
1000
1000
1000

10
10
10
10
10
10

6
6
6
6
6
6

Die Wrmeleitfhigkeit schliesst den Einfluss der Papierdeckschichten ein.

Appendix

226

Tabelle 3 (Fortsetzung)
Stoffgruppe oder Anwendung

Rohdichte

kg/m

Bemessungs- Spezifische
wrmeleit- Wrmespeich
fhigkeit
erkapazitt

cp
W/(mK)
J/(kgK)

Wasserdampfdiffsionswiderstandszahl

trocken

feucht

Erdreich
Ton oder Schlick oder Schlamm
Sand und Kies

1200 1800
1700 2200

1,5
2,0

1670 2500
910 1180

50
50

50
50

2800
2600
1500
1600
2700 3000
2400 2700
2500 2700
2800
2000 2800
1600
1800
2000
2200
2600
2600
400
1750

3,5
2,3
0,85
0,55
3,5
3,5
2,8
3,5
2,2
0,85
1,1
1,4
1,7
2,3
2,3
0,12
1,3

1000
1000
1000
1000
1000
1000
1000
1000
1000
1000
1000
1000
1000
1000
1000
1000
1000

10000
250
30
20
10000
10000
10000
10000
1000
30
40
50
200
250
40
8
50

10000
200
20
15
10000
10000
10000
10000
800
20
25
40
150
200
30
6
40

2000
2100

1,0
1,5

800
1000

40
100

30
60

2300
1000

1,3
0,20

840
1000

10000

10000

500
700

0,13
0,18

1600
1600

50
200

20
50

300
500
700
1000

0,09
0,13
0,17
0,24

1600
1600
1600
1600

150
200
220
250

50
70
90
110

1200

0,23

1500

50

30

Spanplatte

300
600
900

0,10
0,14
0,18

1700
1700
1700

50
50
50

10
15
20

OSB-Platten

650

0,13

1700

50

30

250
400
600
800

0,07
0,10
0,14
0,18

1700
1700
1700
1700

5
10
20
30

2
5
12
20

Gestein
Kristalliner Naturstein
Sediment-Naturstein
Leichter Sediment-Naturstein
Porses Gestein, z.B. Lava
Basalt
Gneis
Granit
Marmor
Schiefer
Kalkstein, extraweich
Kalkstein, weich
Kalkstein, halbhart
Kalkstein, hart
Kalkstein, extrahart
Sandstein (Quarzit)
Naturbims
Kunststein
Dachziegelsteine
Ton
Beton
Platten
Keramik/Porzellan
Kunststoff
(c)
Nutzholz

Holzwerkstoffe
(d)
Sperrholz

Zementgebundene Spanplatte

Holzfaserplatte, einschliesslich MDF

(c)
(d)

(e)

Die Rohdichte von Nutzholz und Holzfaserplattenprodukten ist die Gleichgewichtsdichte bei 20 C und 65% relativer Luftfeuchte.
Als Interimsmassnahme und bis zum Vorliegen hinreichend zuverlssiger Daten knnen fr Hartfaserplatten (solid wood panels,
SWP) und Bauholz mit Furnierschichten (LVL, laminated veneer lumber) die fr Sperrholz angegebenen Werte angewendet werden.
(e) MDF bedeutet Medium Density Fibreboard (mitteldichte Holzfaserplatte), die im sog. Trockenverfahren hergestellt worden ist.
ANMERKUNG 1
Fr Computerberechnungen kann der -Wert durch einen beliebig grossen Wert, wie z. B. 106, ersetzt werden.
ANMERKUNG 2
Wasserdampf-Diffusionswiderstandszahlen sind als Werte nach den in prEN ISO 12572:1999 festgelegten Dry
cup- und Wet cup-Verfahren angegeben.

Appendix

227

Tabelle 4: Wrmeschutztechnische Bemessungswerte fr Mauerwerke, Mrtel und Verputze


Stoffgruppe oder Anwendung

Rohdichte

kg/m
Mauerwerk unverputzt
Modulbackstein Einstein
Modulbackstein Verband
Isolierbackstein
Sichtbackstein
Kaminstein
Kalksandstein

Zementstein
Zementblockstein
Porenbetonstein

Putze, Mrtelschichten
Innenputz fr normale Berechnungen
Aussenputz fr normale Berechnungen
Wrmedmmputz aussen
Wrmedmmputz aussen
Kalkmrtel
Kalkzementmrtel
Zementmrtel
Leichtmrtel

Bemessungs- Spezifische
wrmeleit- Wrmespeich
fhigkeit
erkapazitt

cp
W/(mK)
J/(kgK)

Wasserdampfdiffsionswiderstandszahl

trocken

feucht

1100
1100
1200
1400
1800
1600
1800
2000
2000
1200
300
400
500
600

0,44
0,37
0,47
0,52
0,80
0,80
1,00
1,10
1,10
0,70
0,10
0,13
0,16
0,19

900
900
900
900
900
900
900
900
1000
1000
1000
1000
1000
1000

6
6
6
8
10
25
25
25
15
15
10
10
10
10

4
4
4
6
8
10
10
10
10
10
5
5
5
5

1400
1800
450
300
1800
1900
2200
450
600
900
1600

0,70
0,87
0,14
0,08
0,87
1,00
1,40
0,16
0,21
0,32
0,80

900
1000
1000
1000
1000
1000
1000
1000
1000
1000
1000

10
35
15
15
35
35
35
20
20
20
35

6
15
10
10
15
15
15
5
5
5
15

You might also like