Rug01-002300459 2016 0001 Ac PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 100

Useful energy transfer in air-to-air heat recovery units in

low energy buildings

Willem Faes

Supervisors: Prof. Jelle Laverge, Prof. dr. ir. Michel De Paepe

Counsellors: Prof. dr. ir. Arnold Janssens, Ir. Sven De Schampheleire, Ir. Hugo
Monteyne

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


Master of Science in Electromechanical Engineering

Department of Architecture and Urban Planning


Chair: Prof. dr. ir. Arnold Janssens

Department of Flow, Heat and Combustion Mechanics


Chair: Prof. dr. ir. Jan Vierendeels

Faculty of Engineering and Architecture


Academic year 2015-2016
Useful energy transfer in air-to-air heat recovery units in
low energy buildings

Willem Faes

Supervisors: Prof. Jelle Laverge, Prof. dr. ir. Michel De Paepe

Counsellors: Prof. dr. ir. Arnold Janssens, Ir. Sven De Schampheleire, Ir. Hugo
Monteyne

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


Master of Science in Electromechanical Engineering

Department of Architecture and Urban Planning


Chair: Prof. dr. ir. Arnold Janssens

Department of Flow, Heat and Combustion Mechanics


Chair: Prof. dr. ir. Jan Vierendeels

Faculty of Engineering and Architecture


Academic year 2015-2016
De auteur(s) geeft (geven) de toelating deze masterproef voor consultatie beschikbaar te stellen
en delen van de masterproef te kopieren voor persoonlijk gebruik.
Elk ander gebruik valt onder de bepalingen van het auteursrecht, in het bijzonder met betrekking
tot de verplichting de bron uitdrukkelijk te vermelden bij het aanhalen van resultaten uit deze
masterproef.

The author(s) gives (give) permission to make this master dissertation available for consul-
tation and to copy parts of this master dissertation for personal use.
In the case of any other use, the copyright terms have to be respected, in particular with regard
to the obligation to state expressly the source when quoting results from this master dissertation.

Ghent, June 2016

The author

Willem Faes
Preface

Finishing five years of intensive study is already quite nice, but completing this masters disser-
tation is certainly the frosting on the cake. It was a rather busy year and this thesis demanded
hard work, which would not have been possible without the proper framework I received.

I would first like to thank my supervisors professor Laverge and professor De Paepe, for giving
me the opportunity. Also many thanks to my counsellors Hugo Monteyne, Sven De Scham-
pheleire and professor Janssens.

Throughout the year, in which I learned a lot, the theoretical assistance I got from Hugo
Monteyne and professor Laverge helped me to understand the matter I was working with and
they assisted me in learning how to work with the software. To improve the writing of this
thesis, I received a lot of feedback from Sven De Schampheleire, on both the content as the way
it was written. Finally some parts of the work were done in cooperation with Nils Monsieur,
which contributed to the end result. For all their support, I am very grateful.

Finally I would like to thank my parents and my girlfriend for keeping up with me and always
helping me where possible. Your listening ear and good advice were definitely an added value.

iii
Useful energy transfer in air-to-air heat recovery units in low
energy buildings
Willem Faes

Masters dissertation submitted in order to obtain the academic degree of Master of Science in
Electromechanical Engineering

Supervisors: Prof. Jelle Laverge, Prof. dr. ir. Michel De Paepe


Counsellors: Prof. dr. ir. Arnold Janssens, Ir. Sven De Schampheleire, Ir. Hugo Monteyne

Department of Architecture and Urban Planning


Chair: Prof. dr. ir. Arnold Janssens

Department of Flow, Heat and Combustion Mechanics


Chair: Prof. dr. ir. Jan Vierendeels

Faculty of Engineering and Architecture


Ghent University
Academic year 2015-2016

Summary
The goal of this masters dissertation is to investigate the performance of heat recovery units in
low energy buildings by simulating different buildings under varying conditions. The influence
of several parameters on this performance is studied: the type of building, the insulation and
airtightness, the ventilation flow rates, the ventilation strategy, the heat exchanger effectiveness,
the occupancy pattern and the demanded comfort level. Also a comparison is made between
buildings with energy recovery and buildings with only mechanical exhaust ventilation systems.
The literature review is given in the first part. In its first chapter, ventilation in general is
discussed. The second chapter gives an overview of other research concerning energy recovery
ventilation systems. The literature review is finished with the influence of user behaviour on the
heat demand of a building. The second part of this thesis describes the models of the different
buildings, their ventilation systems and the buildings occupancy. Also the implementation
of these models in the simulation software TRNSYS is discussed here. In the last part, it is
explained how the different performance indicators are defined and the results for the buildings
and the influence of the parameters are discussed.
The results suggest that, although a heat exchanger can have an effectiveness of e.g. 75%,
only approximately 40% of the heat in the extracted air is recovered and supplied usefully to
the rooms. Regarding the type of ventilation system, this study shows that a ventilation system
with energy recovery performs better on heat demand and indoor air quality than a ventilation
system with only mechanical exhaust. However, when the latter is demand-controlled, the heat
demand is similar to that of the ventilation system with energy recovery.

Keywords
Heat recovery ventilation, performance, low energy buildings, simulations

iv
Useful energy transfer in air-to-air heat recovery
units in low energy buildings
Willem Faes
Supervisor(s): Jelle Laverge, Michel De Paepe

Abstract The goal of this paper is to asses the performance of ventila-


tion systems with heat recovery in buildings with a low energy demand for
heating. The amount of heat that is recovered and the fraction of this energy
that is supplied usefully to the building is dependent on several properties
of the building and its ventilation system. The simulation software TRN-
SYS is used to determine the influence of the type of building, its insulation
and level of airtightness, the ventilation flow rates, the ventilation strat-
egy, the effectiveness of the heat exchanger, the occupancy of the building
and the demanded comfort level. Furthermore, a comparison is made be-
tween buildings with a heat recovery ventilation system and buildings with
a mechanical exhaust ventilation system. The results indicate that approx-
imately 40% of the extracted energy is recovered and supplied usefully to
the rooms. A ventilation system with heat recovery performs better than a
mechanical exhaust ventilation system. However when demand-controlled,
the latter has a similar heat demand as a ventilation system with heat re-
covery.
Keywords Heat recovery ventilation, performance, low energy build-
ings, simulations

Fig. 1. Schematic working principle of a fixed-plate heat exchanger [4]


I. I NTRODUCTION
EOPLE spend on average 88% of their time inside a build-
P ing [1]. The indoor air is polluted by different contami-
nants. These pollutants can have a detrimental effect on both
system by simulations in different climatic conditions. The re-
sults are shown in figure 1. The HRV system has a lower energy
the building and its occupants. Some examples of pollutants demand than the MEV system for all climates. However, when
that pose a health risk for people are CO2 , cigarette smoke and the latter is demand-controlled (HCV, humidity controlled venti-
VOCs (volatile organic components) [2]. The integrity of the lation), the energy demand is similar to that of the building with
building can be harmed by e.g. water vapour, which can cause the HRV system. Depending on the climate and the ventilation
the formation of moulds inside the walls [3]. Some of the above flow rates (between brackets), the energy demand is higher or
mentioned contaminants can be avoided (e.g. by not smoking in- lower.
doors), while the sources of other pollutants cannot be avoided, The reduction in heat demand is influenced by several param-
like exhaled CO2 . Typically, this is then solved by replacing eters. For example in the study by El Fouih et al. [4], the ef-
the polluted indoor air with fresh outdoor air with a ventilation fectiveness of the heat exchanger was 60%, so with increasing
system. this value, the heat demand could be further reduced. The influ-
For both the supply and extraction of air to and from the build- ence of the buildings thermal properties on the efficiency of the
ing, two options are possible: naturally or mechanically. A nat- energy recovery was studied by Juodis [5]. The author defined
ural system is wind-driven and the air flows through exhaust or a balance temperature of a building at which the internal heat
supply grilles. With a mechanical system, the air is displaced by gains compensate the heat losses. The closer the external tem-
an electrical fan. A ventilation system where the air is mechan- perature is to this balance temperature, the smaller the efficiency
ically extracted by a ventilator and the supply is wind-driven, of the energy recovery, because the losses are already compen-
is called a mechanical exhaust ventilation system (MEV). With sated by the gains. According to the author, the efficiency of the
a balanced ventilation system, both supply of fresh air and ex- energy recovery decreases when a building has more insulation
traction of polluted air are done with fans, displacing the same and better airtightness, since this reduces the balance tempera-
amount of air. ture of a building.
This last ventilation system allows for a heat exchanger to be On the other hand, in the studies by Binamu and Lindberg [6]
placed between the two air streams. In this case, the ventilation and Roulet et al. [7], it was found that the efficiency of the heat
system is called a heat recovery ventilation system (HRV). The recovery drops when the building has a bad airtightness, because
aim of installing a heat exchanger is to reduce the heat demand no energy can be recovered from air leakages.
of the building by preheating the cold supply air with the energy This paper will determine the efficiency of the energy recov-
in the warmer exhaust air. ery is for different levels of airtightness and the influence of
In a study by El Fouih et al. [4], the heat demand of a building other properties of the building and the HVAC-system (heating,
with a HRV system was compared with a building with a MEV ventilation and air-conditioning) on this efficiency. The param-
eters of which the influence on the heat recovery performance when the maximum CO2 level in the building exceeds 1000 ppm
is studied are: the type of building, the nominal energy demand and is turned back off when this concentration drops below
for heating, the ventilation flow rates, the ventilation strategy, 900 ppm. The atmospheric concentration of CO2 is chosen
the effectiveness of the heat exchanger, the occupancy of the to be 400 ppm and the generation by people is modelled as
building and the demanded comfort level. The model of the 1.2 105 kg/s and 6.7 106 kg/s per person that is awake
buildings and their HVAC-systems and the possibilities for each or asleep respectively [12].
parameter are explained in the next chapter.
For this study, a heating strategy is chosen where not all
II. M ODEL DESCRIPTION rooms are heated simultaneously. A room is instantly heated
A. Building (unlimited heating power) to the desired temperature, only when
people are present in this room. The temperatures to which these
Three types of building are compared in this study. A de- rooms are heated are dependent on the desired comfort level and
tached building is compared with a semi-detached and a ter- are based on the study by Peeters et al. [13]. The authors de-
raced house of similar dimensions and with the same number of scribe correlations for a neutral temperature and the limits of the
rooms. The external wall area is smaller for the semi-detached 10 P P D range (Predicted Percentage Dissatisfied) around this
building and even smaller for the terraced building, because they neutral temperature, as a function of the outdoor temperature.
share one and two walls with an adjacent neighbour respectively. Such correlations are given for the bedrooms, the bathroom and
The detached and semi-detached house are two-storey buildings, other rooms. When heated, the air in the rooms is heated so that
while the terrace house has three storeys. the operative temperature of the rooms equals the desired tem-
Of each type of building, three version are created, differ- perature. The operative temperature is used, because it does not
ing in thickness of insulation and level of airtightness. Each only take the air temperature into account, but also the radiant
type of building has a version with, under a certain set of temperature of nearby surfaces. For example, a person standing
conditions, an energy demand for heating of 15 kW h/m2 a, next to a cold window might have a colder feeling than the actual
30 kW h/m2 a and 60 kW h/m2 a. The version with a heat de- air temperature. The operative temperature is defined by equa-
mand of 15 kW h/m2 a represents a passive house [8]. A low- tion 1, where Tair and Trad are the air and radiant temperature
energy building is modelled by the houses with a heat demand respectively.
of 30 kW h/m2 a [8]. The last version corresponds to most new
houses in the Belgian building stock [9]. To achieve these level Tair + Trad
Top = (1)
of energy demand, first the airtightness was changed for the 2
three versions of the buildings. For the houses with a nomi- Three comfort levels are investigated. Medium comfort is
nal energy demand for heating of 15 kW h/m2 a, 30 kW h/m2 a achieved by heating to the neutral temperatures. For the low-
and 60 kW h/m2 a, an airtightness of respectively 0.6 ach (air est and highest comfort level, the desired temperatures are the
changes per hour), 3 ach and 6 ach are used. Then the thickness lower and upper limit of the 10 P P D range around this neutral
of the insulation in the external walls and the roof was iteratively temperature respectively.
varied until the energy demand for heating reached the desired
value. C. Base case
Most parameters of which the influence is investigated are
B. HVAC-system
now defined. The last variable is the occupancy pattern. The
The heat demand of a building with a HRV system is com- household consists of two members, which are either mostly ab-
pared to the heat demand of a building with an MEV system. sent or mostly at home.
The ventilation flow rates are based on the Belgian standard One base case is used to which the cases with one or more
NBN D 50-001 [10]. The exhaust flow rates are higher for the parameters varied are compared. This base case is the de-
building with the HRV ventilation system, because the total sup- tached building with a nominal energy demand for heating of
ply and exhaust flow rates need to be balanced. To investigate 60 kW h/m2 a. The ventilation is always on, with the nominal
the influence of the ventilation flow rates, the heat demand and flow rates. The heat exchanger has an effectiveness of 75% and
recovery efficiency are also determined when the flow rates are the occupants, who demand a medium comfort level, are mostly
scaled with a factor 2/3 and 1/3. absent.
For the building with the HRV system, the heat exchanger is The buildings are simulated with TRNSYS, which is a simu-
modelled as a plate heat exchanger with a sensible effectiveness lation software used for transient systems [14]. The simulations
of 75%, 80% or 85% at nominal flow rate, based on the effec- are done over one year and one week (to create a balance in the
tivenesses given in the EPB-database [11]. When the ventila- houses). Weather data from Uccle, Belgium, is used to model
tion flow rates decrease, the effectiveness of the heat exchanger the climatic conditions.
increases. This is modelled with the -NTU correlations for a
crossflow heat exchanger. III. P ERFORMANCE INDICATORS
Next to ventilation which is always on, a demand-controlled The evaluation of an energy recovery system is here done
ventilation (DCV) strategy is investigated. This is based on based on five efficiencies, calculated from four heat demands
the maximum CO2 -concentration in the building and is imple- and the (maximum) recovered energy by the heat exchanger.
mented as a simple on/off-strategy. The ventilation is turned on The four different heat demands are:
The actual heat demand, Q, calculated as the integral of the IV. R ESULTS AND DISCUSSION
heating power over the heating season (defined further).
A. Base case
The heat demand without heat exchanger, Q0 , is the heat de-
mand the studied building would have without a heat exchanger Figure 2 shows the resulting heat demands and the (maxi-
or with a heat exchanger with an effectiveness of 0%. mum) recovered energy for the base case.
The heat demand with a perfect heat exchanger, Q1 , is the heat The heat demand, Q, is 30.67 kW h/m2 . The heat exchanger
demand the studied building would have with a (theoretical) heat recovered Qh = 35.57 kW h/m2 of which Q0 Q = 17.74
exchanger with an effectiveness of 100%. was supplied usefully to the rooms. The heat demand could
The heat demand without ventilation, Qnv , is the heat demand even be reduced with Q Q1 = 5.76, kW h/m2 more, with a
the building has when supply and exhaust fans are turned off. perfect heat exchanger. Qnv = 16.83 kW h/m2 gives an indica-
tion of the transmission and infiltration losses. Although for the
Also used for the calculation of the efficiencies are Qh and
calculation of both Q1 and Qnv no energy is released in the at-
Qhm that respectively represent the energy that was recovered
mosphere by the ventilation system, Qnv is still 8.08 kW h/m2
by the heat exchanger and the energy that a perfect heat ex-
lower than Q1 . This is because with a perfect heat exchanger,
changer would have recovered. This last one equals the energy
the colder air from the non-heated rooms is redistributed to the
that is extracted from the rooms by the ventilation system.
heated rooms, increasing the heat demand.
The heating season, over which the heat demands and the
(maximum) recovered energy are calculated, ends the first time The resulting efficiencies are given in figure 3. It can be seen
of the year no heating is required for 24 hours or more. It starts that both 1 and 3 are approximately 75%, which is also the
again after the last time of the year no heating is required for at same as the effectiveness of the heat exchanger. This will be ob-
least 24 hours. served for all other cases as well. From the value of 2 , it can be
seen that almost half of the recovered energy (49.87%) is sup-
The five efficiencies can now be calculated: plied to the building usefully and contributes to the reduction of
Q the heat demand. The other half is supplied to rooms that do not
1 = QQ00Q is the ratio of the reduction in heat demand by
1 require heating at that moment and is then lost through transmis-
the heat exchanger to the maximum heat demand reduction with
sion or infiltration. 4 is 56.17%, which means that 56.17% of
a perfect heat exchanger.
Q0 Q the additional heat required by ventilating the building is recov-
2 = represents the fraction of the recovered energy
Qh ered by the heat exchanger. The last efficiency, 5 = 37.56%,
that contributed to the reduction of the heat demand.
Q shows that of all the heat that the ventilation system extracts
3 = Q h is the ratio of the actually recovered energy to the
hm from the rooms, 37.56% is first recovered and then again sup-
maximum recovered energy.
Q0 Q plied to the building usefully.
4 = Q Q
0 nv
is the fraction of the heat demand caused by
ventilating the building that can be recovered with a heat ex-
changer. B. Influence of parameters
Q Q
5 = Q0 shows how much of the heat extracted by the
To determine the influence of each parameter, the base case is
hm
ventilation system was recovered by the heat exchanger and sup-
plied to the building usefully. used, but each time with one parameter varied. The results are
summarised in figure 4.
On this figure, the bars indicate 5 for each case (left y-axis),
These heat demands and efficiencies are now first illustrated while on the right y-axis, the amount of energy that was saved
for the base case. Then the comparison is made with other cases by installing a heat exchanger, Q0 Q is given. The two hor-
and the influence of the parameters is determined. izontal lines indicate the efficiency and the heat demand reduc-

50 100

40 80 75.49 75.31
Energy, kWh/m2

Efficiency, %

Q nv 60 56.17
30
49.87
Q1
20 Q 40 37.56
Q0
10 Qh 20
Q hm
0 0
21 22 23 24 25

Fig. 2. Heat demand and (maximum) recovered energy for the base case Fig. 3. Efficiencies for the base case
100 30

Heat demand reduction, kWh/m 2


25
Q 0 -Q 25
80
Efficiency, %
20
60
15
40
10

20
5

0 0

y
V
se

t
d

ate

ate

t
%

%
d

for
for
h/m 2

h/m 2

nc
he

ce

DC

80

85
ca

wr

wr

om
om
pa
rra
tac

0=

0=
se

ccu
flo

flo

hc
wc
Te

kW

kW
-de
Ba

2/3

1/3

Hig
ho

Lo
mi

30

15
Se

Hig
Fig. 4. Heat recovery efficiency and heat demand reduction by the heat exchanger for the base case and the variations

tion for the base case. The ventilation flow rate has the biggest
influence on the efficiency, partly caused by the -NTU corre- 40 423
2

361
Heat demand, kWh/m

lations. When the flow rates are scaled with a factor 1/3, the 299 MEV, 15 kWh/m2a
efficiency increases 10.1 percentage points, relative to the base 30 237
175
MEV, 30 kWh/m2a
case. Of course, the indoor air quality (IAQ) deteriorates when MEV, 60 kWh/m2a

the flow rates decrease. Also the occupancy pattern has a high 20 HRV, 15 kWh/m2a
HRV, 30 kWh/m2a
influence on the efficiency, since the efficiency for the case with
HRV, 60 kWh/m2a
a high occupancy is 8.93 percentage points higher than the base 10
case with its low occupancy profile. A higher effectiveness of
the heat exchanger increases the efficiency, but an increase in 0
600 800 1000 1200
effectiveness with 10 percentage points only means an increase CO2-concentration, ppm
in efficiency with 4.21 percentage points.
Fig. 5. Heat demand versus indoor air quality for the MEV and HRV systems
When comparing the heat demand reduction, which e.g. gives in the detached building at different flow rates (in m3 /h)
an indication of the payback time of the investment, it is seen
for the lower flow rates that, although the efficiency increases,
much less energy is recovered (also for the DCV case, since When the MEV building is demand-controlled (DCV), the
this is actually a reduction of the average flow rates). An in- IAQ is even worse, but the heat demand approaches that of the
crease of Q0 Q is observed for the better insulated houses, the HRV building. This is illustrated in figure 6.
cases with a better heat exchanger, the higher occupancy and the The heat demand with DCV is sometimes lower than for the
higher comfort. The type of building does not seem to have a big HRV building at corresponding flow rates. However, when re-
influence on the amount of energy that is recovered usefully. ducing the flow rates of the HRV system, the same IAQ as the
DCV system at high flow rates can be reached with a lower heat
demand.
C. Comparison with MEV systems
When comparing buildings with a heat recovery ventila- V. C ONCLUSION
tion system (HRV) to a mechanical exhaust ventilation system Heat recovery units can recover approximately 40% of the
(MEV), it is important to look at both heat demand and indoor extracted energy and supply it to the rooms to decrease the heat
air quality (IAQ). The results for the detached building at differ- demand. The building itself does not have a great influence on
ent flow rates (total extraction flow rate given between brackets this efficiency, but the ventilation system and the way it is used
in m3 /h) are given in figure 5. Both the heat demand and the do.
IAQ are always better for the HRV buildings. The difference be- Regarding the comparison between HRV and MEV systems,
tween IAQ for corresponding flow rates is the greatest when the HRV systems always perform better regarding IAQ. Buildings
buildings are least airtight. For the better insulated and more air- with an HRV system also have a lower heat demand than build-
tight houses, the difference in heat demand is most significant. ings with an MEV system. This difference in heat demand can
however be eliminated by using a demand-controlled strategy.
40
Heat demand, kWh/m2

DCV, 15 kWh/m2a
30 423 361 DCV, 30 kWh/m2a
299 237
175 DCV, 60 kWh/m2a
HRV, 15 kWh/m2a
20
HRV, 30 kWh/m2a
HRV, 60 kWh/m2a
10

0
600 800 1000 1200
CO2-concentration, ppm

Fig. 6. Heat demand versus indoor air quality for the DCV and HRV systems
in the detached building at different flow rates (in m3 /h)

R EFERENCES
[1] AP Jones. Indoor air quality and health. Atmospheric environment,
33(28):4535-4564, 1999.
[2] World Health Organization et al. Air quality guidelines for Europe, 2000.
[3] L Molhave, B Bach, and OF Pedersen. Human reactions to low concentra-
tions of volatile organic compounds. Environment International, 12(1):167-
175, 1986.
[4] YE Fouih, P Stabat, P Riviere, P Hoang and V Archambault, Adequacy of
air-to-air heat recovery ventilation system applied in low energy buildings,
Energy and Buildings, 54:29-39, 2012.
[5] E Juodis. Extracted ventilation air heat recovery efficiency as a function
of a buildings thermal properties, Energy and Buildings, 38(6):568-573,
2006.
[6] A Binamu and R Lindberg. The impact of air tightness of the building
envelope on the efficiency of ventilation systems with heat recovery. 2001.
[7] CA Roulet, FD Heidt, F Foradini and MC Pibiri. Real heat recovery with
air handling units. Energy and Buildings, 33(5):495-502, 2001.
[8] Vlaanderen. Belastingvermindering voor lage-energiewoningen, passief-
woningen en nul-energiewoningen. https://www.vlaanderen.be/nl/bouwen-
wonen-en-energie/bouwen-en-verbouwen/belastingvermindering-voor-
lage-energiewoningen-passiefwoningen-en-nul-energiewoningen,. Ac-
cessed: 30-04-2016.
[9] Vlaams Energieagentschap. Cijferrapport: 10 jaar energieprestatieregelgev-
ing. Procedures, resultaten en energetische karakteristieken van het Vlaams
gebouwenbestand - periode 2006-2015. 2016.
[10] BIN. NBN D 50-001: Ventilatievoorzieningen in woongebouwen, 1991.
[11] EPB-productgegevens databank - V ENTILATOR EN V ENTILATIEGROEP -
Erkende productgegevens (status 1), 2016.
[12] ASHRAE Fundamentals Handbook. American Society of Heating, Re-
frigerating and Air-conditioning Engineers. Inc.: Atlanta, GA, USA, 2009.
[13] L Peeters, R de Dear, J Hensen and W Dhaeseleer. Thermal comfort in
residential buildings: Comfort values and scales for building energy simu-
lation. Applied Energy, 86(5):772-780, 2009.
[14] S.A. Klein et al. Transient System Simulation Program (TRNSYS 17)
Manuel. Thermal Energy System Specialists, Madison, USA, 2010.
Contents

Preface iii

Abstract iv

Extended abstract v

Nomenclature xiv

1 Introduction 1
1.1 The big picture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Goal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

I Literature review 4

2 Ventilation 5
2.1 Why ventilate? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.2 Ventilation systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.3 Ventilation flow rates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.4 Demand-controlled ventilation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

3 Energy recovery ventilation 10


3.1 Energy recovery technologies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3.1.1 Fixed-plate heat exchanger . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3.1.2 Rotary wheel heat exchanger . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.1.3 Heat pipe recovery units . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.1.4 Run-around coils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.2 Ventilation heat load reduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.2.1 Thermal properties of the building . . . . . . . . . . . . . . . . . . . . . . 15
3.2.2 Climate dependency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.2.3 Frosting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

4 Human influences 20
4.1 User behaviour . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4.2 Thermal comfort . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4.2.1 Standard approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4.2.2 Alternative approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

xi
CONTENTS xii

II Model description 30

5 Building 31
5.1 Building types . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
5.2 Energy demand for heating . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
5.2.1 Test conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
5.2.2 Insulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
5.2.3 Infiltration, exfiltration and internal air flows . . . . . . . . . . . . . . . . 33

6 HVAC-system 34
6.1 Ventilation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
6.1.1 Flow rates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
6.1.2 Flow controllers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
6.1.3 Heat exchanger . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
6.1.4 Demand-controlled ventilation . . . . . . . . . . . . . . . . . . . . . . . . 36
6.2 Heating . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

7 Internal heat gains and occupancy 37


7.1 Internal heat gains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
7.2 Occupancy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

8 Parameter summary 40

III Results and discussion 42

9 Heat recovery performance 43


9.1 Heat demand . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
9.2 Heating season . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
9.3 Recovered energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
9.4 Efficiency definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

10 Base case 47

11 Influence of parameters on recovery performance 51


11.1 Building type . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
11.2 Nominal energy demand for heating . . . . . . . . . . . . . . . . . . . . . . . . . 52
11.3 Ventilation flow rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
11.4 Demand-controlled ventilation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
11.5 Heat exchanger effectiveness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
11.6 Heating and occupancy pattern . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
11.7 Comfort level . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
11.8 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

12 Comparison with type C ventilation systems 60


12.1 No demand-controlled ventilation . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
12.2 Demand-controlled ventilation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

13 Conclusion 65

Bibliography 67

Appendices 72
CONTENTS xiii

A Floor plans 73

B Building composition 77

C Calculation of the air mass flow coefficient for crack modelling 79

D TRNSYS model 81
Nomenclature

Abbreviations

ach Air changes per hour


DCV Demand-controlled ventilation
ERV Energy recovery ventilation
HDD Heating degree day
HRV Heat recovery ventilation
HVAC Heating, ventilation and air conditioning
IAQ Indoor air quality
NTU Number of transfer units []
PMV Predicted mean vote []
PPD Predicted percentage dissatisfied []
VOC Volatile organic component

Symbols

A Surface area [m2 ]


C Heat capacity rate [J/sK]
C Heat capacity rate ratio []
Cq Air volume flow coefficient in flow controller model [m3 /h @ p = 1 P a]
Cs Air mass flow coefficient in crack model [kg/s @ p = 1 P a]
H Specific enthalpy [J/kg]
m Air mass flow rate [kg/s]
n Air mass flow exponent []
n50 Airtightness [ach]
P Heating power [kW ]
p Pressure difference [P a]
Q Heat demand [kW h/m2 ]
T Temperature [ C]
U Heat transfer coefficient [W/m2 K]
V Volume [m3 ]
V Volume flow rate [m3 /h]

Greek symbols

Effectiveness []
Efficiency []
Density [kg/m3 ]
Humidity ratio [kg/kg]

xiv
NOMENCLATURE xv

Subscripts

0 No heat exchanger used


1 With a perfect heat exchanger used
50 Determined with a pressure difference of 50 P a
e Exhaust
e,ref External reference
f Fresh
H Enthalpy
h Recovered
hm Maximally recoverable
L Latent
max Maximum
min Minimum
n Neutral
nv No ventilation
r Roof
s Sensible
t Total
w Wall
Chapter 1

Introduction

1.1 The big picture


One of the Europe 2020 targets is to have more sustainable energy and change the climate for the
better [1]. To do so, the EU wants to cut greenhouse gas emissions with 20%, increase the share
of renewable energy in the final energy consumption to 20% and have 20% more energy efficiency.

The primary energy consumption in Europe from 1990 to 2013 is shown in Figure 1.1, as
well as the projected path to the 2020 target. Progress has been made since 2006, attributed to
the economical crisis, but also to energy efficiency policies, reducing the energy use of transport,
buildings and industry. Transport is still the mayor energy consuming sector, closely followed
by the households, as can be seen from Figure 1.2.

Figure 1.1: Primary energy consumption in million tonnes of oil equivalent (Mtoe) in Europe, 1990-
2013. Source: Eurostat

With the need for an increase of the energy efficiency in mind, the EU has created energy
labels for appliances, where the most energy efficient and ecologic products are labelled A+++
[2]. Since households, as can be seen on Figure 1.2, represent 26.8% of the primary energy con-
sumptions, it is important to build new houses with a heat demand as low as possible. By 2020
all new buildings need to be nearly zero-energy and energy performance certificates, similar to
the labelling of appliances, need to be provided when purchasing or renting a house.

To improve the energy performance of houses, more insulation is used during the construction
or renovation and the airtightness of the buildings is improved, reducing infiltration of cold air in
the building. This reduction of air leakages has the bad consequence that the indoor air, which

1
CHAPTER 1. INTRODUCTION 2

Figure 1.2: Final energy consumption in Europe, 2013. Source: Eurostat

is polluted by e.g. CO2 produced by people, is less frequently replaced by fresh outdoor air. The
problem of a bad indoor air quality (IAQ) can be overcome with the use of ventilation, either
naturally or mechanically. Natural ventilation systems are wind driven, while a mechanical
ventilation system uses fans to supply, extract or supply and extract air.
A heat exchanger can be used when a building has both mechanical supply of fresh air and
mechanical extraction of polluted indoor air. Such a device is then installed between the two air
streams and preheats the fresh air entering the building. Since the air supplied to the building
is warmer than without a heat exchanger, the heat demand of the building is reduced.

1.2 Goal
The performance of buildings equipped with an energy recovery system is determined in this
thesis. To do so, simulations are done with the software package TRNSYS and the efficiency of
the energy recovery and the heat demand reduction are calculated.

The influence of several parameters on the performance is studied. These parameters are:
the type of building, the insulation and airtightness of the building envelope, the ventilation flow
rates, the ventilation strategy, the effectiveness of the heat exchanger, the desired comfort level
and the level of occupancy. One base case is defined to which the other cases are compared.
Furthermore, a comparison is made with buildings with only mechanical extraction. The
relation between the ventilation flow rates, the energy demand and the indoor air quality is
investigated.

Juodis [3] already illustrated that for more insulated and airtight buildings, the efficiency of
the energy recovery is lower, since the relative importance of the internal heat gains increases.
On the other hand, it was shown by both Binamu and Lindberg [4] and Roulet et al. [5] that
with increasing amounts of infiltration, caused by a low airtightness, the possible energy recovery
decreases.
CHAPTER 1. INTRODUCTION 3

Regarding the comparison between buildings with energy recovery and buildings with only
mechanical extraction, El Fouih et al. [6] did simulations for different climatic conditions. The
annual primary heat demand was lower for buildings with an energy recovery ventilation system
for all climates. However, when the ventilation in the building with mechanical extraction was
demand-controlled, the energy demand decreased to approximately the same level as the building
with energy recovery.

1.3 Outline
This thesis consists of three parts.
In the first part, the literature review is given, discussing first ventilation in general. Then an
overview of the existing research on energy recovery ventilation is given and finally the influence
of user behaviour on the heat demand is discussed.
The used model is described in the second part. First the buildings itself are handled. Next the
HVAC-system (heating, ventilation and air condition) is described. Also the internal heat gains
and the occupancy of the building are explained here.
The third and last part of this masters dissertation defines the performance indicators and
gives the results. The base case, the influence of the parameters and the comparison with the
ventilation system with natural supply are discussed consecutively.
Part I

Literature review

4
Chapter 2

Ventilation

In this chapter, ventilation will be discussed generally. First, why a building should be ventilated
and how this can be done is explained. Finally, the required ventilation flow rates and different
control strategies will be given.

2.1 Why ventilate?


People spend on average 88% of their time inside a building [7]. The indoor air is polluted by
different contaminants:

Carbon dioxide (CO2 ) naturally occurs in the atmosphere. The current concentration
of CO2 outdoors is about 385 ppm according to Hansen et al. [8]. Indoors, people and
combustion appliances produce CO2 and elevate the concentration with approximately
700 ppm. 1000 ppm to 1200 ppm is an acceptable level of indoor CO2 -concentration, where
concentrations of 5000 ppm or more form a serious health risk [9].

Incomplete combustion from combustion appliances and vehicles as well as the smoking
of cigarettes produces carbon monoxide (CO), a toxic odourless gas. The global back-
ground concentration of CO ranges between 0.05 ppm and 0.12 ppm [10]. When inhaled,
CO forms COHb with haemoglobin in the blood and impairs the oxygen-carrying capacity
of the blood. The concentration of COHb in the blood is dependent on the concentration
of CO in the air and the duration of the time spent is this air. The air quality guide-
lines for Europe by the WHO [10] recommend maximum CO-concentrations of 87 ppm for
15 min, 52 ppm for 30 min, 26 ppm for 1 h and 9 ppm for 8 h. In 2014, 22 people died due
to CO-intoxication in Belgium [11].

Several minerals used in building materials contain radium-226, which decays to radon,
an inert radioactive gas. Radon itself does little harm, but it decays and produces ra-
dioisotopes called radon daughters or progeny [7]. These progeny are electrically charged,
and when inhaled, they remain in the lungs and can cause cancer. The radon concentra-
tion (usually expressed in becquerel per cubic meter) of the outdoor air is approximately
10 Bq/m3 [10]. In the WHO guidelines, no guideline value for radon concentration is
recommended, but the concentration should be kept as low as possible to minimize lung
cancer risk.

All molecules containing at least one carbon and one hydrogen atom are called organic
compounds. Some of these compounds have a low boiling point (between 50 C and 250 C)
and off-gas vapours into the air [12]. These compounds are called volatile organic com-
pounds (VOCs). Many products use materials containing VOCs because of their good
properties as, e.g.: insulating, relatively cheap and fire-resistant [7]. Common sources

5
CHAPTER 2. VENTILATION 6

of VOCs are: consumer product, paints, building materials, combustion appliances and
clothing. Concentrations of VOCs in buildings are typically expressed with the level of
total volatile organic compounds (TVOCs), instead of using the concentration of indi-
vidual VOCs. According to Jones [7], mean values of the concentration range between
200 500 g m3 . It should be noted that people are exposed to slightly higher values,
since they are often close to a source (e.g. clothing). The effects of exposure to VOCs was
described by [13]. Health problems caused by VOCs can be both acute and chronic. Usu-
ally, low concentrations of VOCs lead to respiratory problems (especially with asthmatics),
dryness of the nose and irritation of the eyes.

Several other pollutants are found in indoor air, most of them with negative impact on
either the building or the people inside. Some examples are: Water vapour, cigarette
smoke, dust, ozone, asbestos,...

Some of the pollutants discussed above can be avoided (e.g. by not smoking in the house).
When the sources of pollutants cannot be eliminated (e.g. exhaled CO2 ), the contaminated air
has to be replaced by fresh and clean air to guaranty human health, productivity and comfort.

2.2 Ventilation systems


The purpose of a ventilation system is to extract indoor air, polluted with contaminants as
described in Section 2.1, to supply fresh outdoor air and to properly distribute the air in the
house [14]. A good ventilation system maintains a good indoor air quality with minimal energy
losses. Usually, fresh air is supplied in dry rooms (e.g. living, bedrooms and office) and moist
or contaminated air extracted from wet rooms (e.g. toilet, bathroom and kitchen). The air is
transported through cracks and interspaces (e.g. hall).
Since extraction and supply of air can be performed both naturally or mechanically, four
types of ventilation systems (A, B, C and D ventilation) are described by the Belgian standard
NBN D 50-001 [15]. A ventilation system is called type A when exhaust as well as supply are
done naturally. Fresh air is supplied through supply grilles and discharged through exhaust
grilles. The flow from dry to wet rooms occurs through grilles in doors and walls and through
slits under doors without the use of mechanical power (Figure 2.1). System A ventilation is
cheap to install and requires little maintenance. Since ventilation is dependent on wind and air
pressure it is difficult to regulate. High wind speeds usually include high energy losses, while
for low wind speeds, ventilation might be insufficient [16].

Figure 2.1: System A ventilation

In the past, when the insulation and airtightness of the buildings was low, system A venti-
lation was suited and commonly used. Because of the high energy losses, there is the tendency
to build more airtight. This creates the need for mechanical ventilation systems.
System B ventilation consists of natural extraction and mechanical supply. Fresh air enters
the building with electrical ventilators in dry rooms (Figure 2.2). As opposed to system A
CHAPTER 2. VENTILATION 7

ventilation, the necessary amount of fresh air can be supplied for all wind speeds with system
B ventilation. Some disadvantages are the electric power consumption of the fan, the con-
stant overpressure and the inability to control the extraction of pollutants. Because of these
disadvantages, system B ventilation rarely occurs in practice.

Figure 2.2: System B ventilation

A better control of the extraction of pollutants can be achieved with system C ventilation
(mechanical extraction and natural supply, see Figure 2.3). This type of ventilation occurs
frequently in residential buildings with moderate airtightness. The power use of the fan can be
limited by a demand-controlled ventilation strategy, where air is only extracted when necessary.
Such systems are commercially often called ventilation system C+. With system C ventilation,
the room is always at a pressure below atmospheric pressure. This creates the danger of flue
gasses entering the house instead of leaving through the chimney. With system C ventilation,
thermal recovery of the energy in the exhaust air stream is possible through an exhaust air heat
pump [17].

Figure 2.3: System C ventilation

Mechanical extraction in combination with mechanical supply is called system D ventilation


(Figure 2.4). The term balanced ventilation is also used. By balancing the supply and extraction
flow rates, the risk of creating a pressure lower or higher than atmospheric pressure is reduced.
However, system D is more difficult to install in comparison with system C and B and the
operation of two fans requires a higher electrical power.
System D ventilation is commonly used in combination with energy recuperation. By in-
stalling a heat exchanger between inlet and outlet streams, sensible and latent heat can be
transferred between fresh and exhaust air. The energy recovery possibilities are discussed in
more detail in Chapter 3.

2.3 Ventilation flow rates


The necessary ventilation flow rates for Belgian houses are prescribed by the standard NBN D
50-001: Ventilation systems for housings [15]. The required ventilation flow rates are given in
Table 2.1. In this table, the first number dictates the nominal flow rate. In the norm, minimal
CHAPTER 2. VENTILATION 8

Figure 2.4: System D ventilation

flow rates and flow rates to which the ventilation flow rate may be limited are given. These
values are given between brackets. For example, in the living room a nominal flow rate of
3.6 m3 /h m2 is required. A minimal flow rate of 75 m3 /h is necessary, but it may be limited to
150 m3 /h if the floor surface area times 3.6 were to exceed 150.

Table 2.1: Ventilation flow rates for different rooms [15]

Room Ventilation flow rate


Living room 3.6 m3 /h m2 (75 m3 /h - 150 m3 /h)
Bedroom, study room and play room 3.6 m3 /h m2 (25 m3 /h - 72 m3 /h)
Kitchen, bathroom and storage 3.6 m3 /h m2 (50 m3 /h - 75 m3 /h)
Open kitchen 3.6 m3 /h m2 (75 m3 /h - /)
Toilet 25 m3 /h
Hall 3.6 m3 /h m2

The flow openings, serving as supply and exhaust openings for the wet and dry rooms
respectively, should allow sufficient air flow. The necessary flow rate at P = 2 P a are tabulated
in Table 2.2. This table also shows the minimum surface area of the openings, if they consist of
a non-closeable slit under the door.
Table 2.2: Flow opening requirements @ p = 2 P a [15]

Room Flow rate Surface area


Dry rooms: Living room,
25 m3 /h 70 cm2
bedroom, study room and play room
Wet rooms: Bathroom,
25 m3 /h 70 cm2
storage and toilet
Kitchen 50 m3 /h 140 cm2

2.4 Demand-controlled ventilation


The ventilation flow rates prescribed by the standard are not optimal in every situation. For
example, when nobody is present in the house, ventilation could be lowered or turned off. This
would lower the energy demand, while maintaining sufficient indoor air quality. On the other
hand, when a lot of people are present, CO2 concentration could locally rise above recommended
values. In this case, higher ventilation flow rates could be desired. The ventilation strategy
controlling the ventilation based upon the occupancy is called demand-controlled ventilation
(DCV). Although mostly the concentration of CO2 is used to determine the occupancy level [18]
CHAPTER 2. VENTILATION 9

and adapt flow rates accordingly, other strategies, e.g. based on water vapour [19; 20] (humidity
controlled ventilation), are possible as well.
Because of demand-controlled ventilation, a lower electric power is required for the fans.
Combined with the reduced heating or cooling accompanied with ventilation, significant energy
savings are possible. In a study with both simulations and experiments by Lu et al. [21], savings
up to 34% of the energy related to ventilation are reported. Brandemuehl and Braun [22]
studied DCV for different climates and schedules of occupancy. In all cases, energy savings
were significant. The authors note that savings are highest for buildings with large variability
in occupancy. The savings in cold climates (heating) are more significant than in hot climates
(cooling), but still are very dependent on occupancy.
Some risk inherent to DCV should be considered. The first problem are the non-occupant
related contaminants that can accumulate at low occupancy level. According to Chao and Hu
[23], using only DCV, these pollutants (like radon and VOCs) might reach unacceptable levels.
Next, enough CO2 -sensors should be installed, because a single sensor in the common exhaust
of a multi-zone building can cause under-ventilation in the occupied rooms, as described by
Mumma [24]. This is why Schell et al. [25] recommend, based on experience, a base ventilation
rate of 15% to 50% of the design ventilation rate. Newly build houses require a higher base
ventilation rate than older buildings to account for the off-gassing of building components.
Chapter 3

Energy recovery ventilation

As mentioned higher, the possibility for energy recovery exists with mechanical ventilation (sys-
tem D). In this chapter, the different technologies for energy recovery are first discussed. Next,
an overview is given of other aspects that influence the energy recovery potential.

3.1 Energy recovery technologies


Different devices for energy recovery have been developed. The four technologies that are de-
scribed are: fixed-plate heat exchanger, rotary wheel energy exchanger, heat pipe recovery units
and run-around coils.
The energy transfer possibilities of a heat exchanger are expressed by the effectiveness. A sen-
sible, latent and enthalpy (or total) effectiveness can be defined and are respectively given in
(3.1), (3.2) and (3.3).
me (Tei Teo ) mf (Tf i Tf o )
S = = (3.1)
mmin (Tei Tf i ) mmin (Tei Tf i )
me (ei eo ) mf (f i f o )
L = = (3.2)
mmin (ei f i ) mmin (ei f i )
me (Hei Heo ) mf (Hf i Hf o )
H = = (3.3)
mmin (Hei Hf i ) mmin (Hei Hf i )
In these equations subscripts i and o refer to inlet and outlet, while subscripts e and f refer to
exhaust and fresh air. mmin is defined as min(me , mf ).

3.1.1 Fixed-plate heat exchanger


According to Mardiana-Idayu and Riffat [26], the most common type of air-to-air heat exchangers
for energy recovery ventilation is the fixed-plate heat exchanger. Several internal air streams
are created and separated by thin plates. Energy is transferred from the hot air stream to the
cold air stream. The working principle is shown in Figure 3.1. Because the two air flows are
separated by plates, the risk for cross contamination (mixing of the two air flows) is low. The
plates can be corrugated to improve heat transfer.
Different possible airflow configurations are parallel flow, counter flow and cross flow. Fixed-plate
heat exchangers can have a high effectiveness. Effectivenesses up to 94% in optimal conditions
have been reported by manufacturers [27].
Vapour permeable membranes can be used in the construction of the plates, as described by
Sander and Janssen [28]. This allows vapour transfer from a humid air stream to a dry air
stream. Nasif et al. [29] tested 60 gsm Kraft paper as a permeable membrane. Their study
showed a sensible effectiveness up to 75% and a latent effectiveness up to 60%.

10
CHAPTER 3. ENERGY RECOVERY VENTILATION 11

Figure 3.1: Schematic working principle of a fixed-plate heat exchanger [26]

The effectiveness of a fixed-plate heat exchanger is dependent of the operating conditions.


In a study by Min and Su [30] the influence of outdoor temperature and humidity in both
hot and cold weather on the effectiveness of a membrane-based fixed-plate heat exchanger is
investigated with a numerical model. In hot and humid weather, where heat and moisture are
transferred from supply to exhaust air, the sensible effectiveness shows little change with outdoor
temperature, but a small decrease with increasing humidity. The latent effectiveness decreases
slightly with increasing outdoor temperature and increases more sharply with increasing outdoor
humidity. This is illustrated in Figure 3.2. In cold and dry weather (only temperatures above

Figure 3.2: Influence of outside humidity on sensible, latent and enthalpy effectiveness for different
outside temperatures in hot weather [30]

the freezing point of water are investigated, to avoid problems with frosting, as described in
Section 3.2.3), where heat and moisture are transferred from exhaust to supply air, the sensible
effectiveness shows little change with both varying outdoor temperature and humidity. As seen in
Figure 3.3, the latent effectiveness decreases with increasing outdoor temperature, but increases
with increasing outdoor humidity. Furthermore, the sensible effectiveness shows little difference
between hot and cold weather, but the latent effectiveness is significantly smaller in hot weather
than in cold weather. The effectiveness changes are attributed to the thermal resistance of
the membrane and the resistance of the membrane to moisture transfer that are influenced by
outdoor temperature and humidity.
Niu and Zhang [31] came to the conclusion that sensible effectiveness is not depending on
CHAPTER 3. ENERGY RECOVERY VENTILATION 12

Figure 3.3: Influence of outside humidity on sensible, latent and enthalpy effectiveness for different
outside temperatures in cold weather [30]

outdoor state, but latent effectiveness is strongly coupled with operating conditions. This is
illustrated for outdoor temperature in Figure 3.4.

Figure 3.4: Influence of outside temperature on sensible, latent and enthalpy effectiveness [31]

The effect of the air flow rate on the effectiveness of the heat exchanger can by analysed with
the -NTU method [32]. The number of transfer units (NTU) designates the non-dimensional
heat transfer size of the heat exchanger and is defined as:
AU 1
Z
NTU = = U dA (3.4)
Cmin Cmin A
with A the heat transfer surface area, U the average heat transfer coefficient and C the capacity
rate (C = m cp ). Cmin and Cmax are the capacity rates of the air flow (supply or exhaust) with,
respectively, the lowest and the highest capacity rate. It can be shown that the effectiveness
is a function of N T U , the capacity rate ratio (C = Cmin /Cmax ) and the flow configuration.
Equations (3.5) and (3.6) give the -NTU expression for a counterflow heat exchanger. Sim-
ilar correlations exist for other flow configurations.
1exp[N T U (1C )]
= 1C exp[N T U (1C )] , C 6= 1 (3.5)
= 1+N NT U
TU , C = 1 (3.6)
CHAPTER 3. ENERGY RECOVERY VENTILATION 13

This relationship is shown for different values of C in Figure 3.5. When the mass flow rate and
thus Cmin increases (or N T Umax decreases), the effectiveness drops.

100

80
Effectiveness 0, %

60

40
C* = 0
C* = 0.25
20 C* = 0.5
C* = 0.75
C* = 1
0
0 1 2 3 4 5
Number of transfer units, NTU = UA/C
max min

Figure 3.5: Sensible effectiveness of a counterflow heat exchanger for different mass flow rates

3.1.2 Rotary wheel heat exchanger


The rotary wheel recovery type (energy wheel), visualised in Figure 3.6, consists of a storage
mass in a casing rotating between the fresh air stream and the exhaust air stream [26]. The rotor
is driven by an electrical motor, usually rotating in a range of 3 to 15 rpm. The storage mass,
usually a desiccant coated matrix, provides the possibility of moisture transfer. Disadvantages
of the rotary wheel are the required electrical power to run the wheel and the greater risk for
cross contamination caused by leaks and carry-over.

Figure 3.6: The rotary wheel energy recovery type [26]

Zhang and Niu [33] studied the effectiveness of a rotary wheel heat exchanger. The effective-
ness is influenced by the material and thickness of the storage mass and has a maximum for a
certain rotational speed. For a rotary heat exchanger consisting of a honeycomb structure with
a wall thickness of 0.2 mm, the effectiveness for various rotational speeds is given in Figure 3.7.
An optimal rotational speed is observed. Zhang and Niu [33] explain that on the one hand for
higher rotational speeds, the time for adsorption and regeneration is to short. On the other
hand, for lower rotational speeds, the adsorption process is less effective. In their study, the
CHAPTER 3. ENERGY RECOVERY VENTILATION 14

optimal rotational speed is 10 rpm for both sensible and latent effectiveness. a On the website

Figure 3.7: Recovery performance at various rotational speeds [33]

of Eurovent (the European Committee of Air Handling and Refrigeration) [34], effectivenesses of
energy wheels by different manufacturers are listed. Typical sensible and latent effectivenesses
are 80% and 60% respectively. Although the effectiveness is also depending on the model and
the manufacturer, it can be observed that, when increasing the air speed from 1.5 m/s to 4 m/s,
the efficiency can drop more than 10 percentage points.

3.1.3 Heat pipe recovery units


Heat transfer in ventilation systems can also been done by heat pipes [26]. By using the vaporisa-
tion heat of the working fluid, heat can be transferred, even with small temperature differences.
In Figure 3.8, the working principle of a heat pipe recovery system is illustrated. Heat pipe
recovery units have a low pressure drop, but moisture transfer is impossible. The two air flows
need to be brought close together (short length of the heat pipes), but can remain separated
(low risk for cross contamination) [35]. With heat pipe recovery units, there is no need for other
pumps or drives. Reported effectivenesses of heat pipe recovery units are between 45% and 50%
[26].

Figure 3.8: Working principle of a heat pipe recovery system [26]


CHAPTER 3. ENERGY RECOVERY VENTILATION 15

3.1.4 Run-around coils


When the air supply and exhaust ducts cannot be placed next to each other, heat transfer by a
run-around heat recovery system is possible [26]. As seen in Figure 3.9, a third fluid, like e.g.
water, is pumped between the two air streams to transfer the heat. Because of the physical
separation of the ducts there is a high flexibility and low risk for cross contamination. Run-
around coils have an effectiveness between 45% and 65% and have the disadvantage of requiring
a pump for the third fluid [26].

Figure 3.9: Schematic working principle of a HRV-system with a run-around [26]

3.2 Ventilation heat load reduction


From here on, the heat load imposed on the heating system by both natural infiltration and
mechanical ventilation is called the ventilation heat load (based on Fisk and Turiel [36]). The
aim of installing an energy recovery system between supply and exhaust flows is to reduce this
heat load. In a study by El Fouih et al. [6], a residential building was simulated in cities with a
different climate in France to compare type C, C+ and D ventilation systems. Their results are
given in Figure 3.10. For all climates, the energy demand was lower for the type D ventilation
system (HRV, heat demand ventilation) than for the type C building (MEV, mechanical exhaust
ventilation). When comparing the type D and type C+ (HCV, humidity controlled ventilation),
it can be seen that the HCV scores better when the flow rates (indicated between brackets) are
lowest. Of course, this means that the indoor air quality will be worse.
The reduction in heat demand is influenced by several parameters. For example, in the study
by El Fouih et al. [6], the effectiveness of the heat exchanger was 60%, so by increasing this
value, the heat demand could further be reduced. Not only the heat exchanger effectiveness is a
determining factor of the magnitude of this reduction. Other aspects that play a significant role
in the operation and efficiency of the energy recovery unit are the buildings thermal properties,
the climate it is situated in, the possibility and effects of frosting and the way the occupants
use the house. The first three aspects are described in this section. Chapter 4 deals with the
human influences on the heat demand.

3.2.1 Thermal properties of the building


The heat losses of a building are strongly determined by the insulation of its external walls and
the roof (transmission losses) and the airtightness of the building envelope (infiltration losses).
Since the internal heat gains are not influenced by the thermal properties of the building, the
insulation and airtightness also have an influence on the heat gain/loss ratio of the building and
CHAPTER 3. ENERGY RECOVERY VENTILATION 16

Figure 3.10: Primary energy consumption for the three ventilation system in the house in different
climates [6]

its balance temperature. The balance temperature is this external temperature at which the
losses equal the internal heat gains and no additional heat is required [37]. According to Juodis
[3], the efficiency of the energy recovery decreases when the external temperature approaches
the balance temperature of the building, because the losses are compensated by the gains and
the recovered energy no longer contributes to the heating of the house.

Although a higher level of airtightness means a lower balance temperature so less efficient
recovery, other authors say that a good level of airtightness is important [4][5], because the goal
of energy recovery ventilation is to release the least amount of energy in the atmosphere. This is
not possible when great volumes of air leak through cracks in the envelope of the building. This
infiltration and exfiltration often occurs in garages and attics, which respectively represent, on
average, 30% and 48% of the total leakage, according to Laverge et al. [38]. The airtightness of a
building is measured by performing a pressurization test (also called blowerdoor test). A pressure
difference (50 P a for Flemish EPB-regulations) is created between indoor and outdoor and the
air flow rate displaced by the fan, V50 , is measured. The n50 -value, (n50 = V50 /V ), expresses
the airtightness of the building, where V is the total volume of the building. Airtightness is
expressed in ach (air changes per hour).
The impact of airtightness on heat recovery ventilation was studied by Binamu and Lindberg
[4]. The possible energy recovery decreases significantly with decreasing airtightness, as seen in
Figure 3.11. The authors suggest that for a building equipped with heat recovery ventilation,
the airtightness should not exceed 1.5 ach at 50 P a.
The same conclusion is found by Roulet et al. [5]. The authors calculate a global heat
recovery efficiency based on heat exchanger effectiveness, infiltration and exfiltration rates and
recirculation ratios (part of the air that is recirculated). A drop in heat recovery efficiency was
reported from efficiencies ranging between 50% to 90% to efficiencies ranging between 5% to
68% when incorporating leaks.

3.2.2 Climate dependency


In the United States, an analytical study by Fisk and Turiel [36] was conducted in four cities
with different climates. The heat load of a typical house (infiltration rate of 0.60 ach) was
compared with the heat load of a tight house with heat exchanger (infiltration rate of 0.2 ach)
CHAPTER 3. ENERGY RECOVERY VENTILATION 17

Figure 3.11: Comparison of energy losses associated with ventilation between a fully airtight building
and a leaky building [4]

in two situations (different ventilation rate and heat exchanger effectiveness). In the situation
with high ventilation rate and high heat exchanger effectiveness, they found a ventilation heat
load reduction in Minneapolis (4657 HDD) of 11.9 GJ or 45% and 5.3 GJ or 53% in Atlanta
(1645 HDD).
A more recent study by Rasouli et al. [39] investigated possible savings on heating and
cooling load in four different cities in the United States in a 10-storey building. The study using
TRNSYS compared different sensible and latent effectiveness values. It was found that energy
savings with an energy recovery ventilator (l 6= 0) can be 15% higher than with sensible-only
heat exchangers (l = 0). This was especially the case in humid climates, because there, an
energy recovery ventilator reduces the dehumidification load significantly. They also concluded
that the impact of the sensible effectiveness is most significant in the coldest climate since there,
the sensible heat load accounts to 90% of the total heat load.

3.2.3 Frosting
When outdoor temperatures are low, ice or frost formation inside the heat exchanger is possible
[40]. This frosting has an adverse effect on the pressure drop and causes a higher thermal
resistance. Emery and Siegel [41] report a reduction in heat transfer from 50% to 75%.
The physical process of frost formation is described by Rafati Nasr et al. [42]. For a heat
exchanger without moisture transfer (only sensible energy transfer), condense appears on the
surface when the exhaust air is cooled below its dew temperature. When the surface temperature
falls below the freezing temperature of water, the condensed water freezes and frost is formed
(see Figure 3.12). The process for a heat exchanger with moisture transfer is illustrated on the
psychrometric diagram of Figure 3.13. The possibility of moisture transfer causes the temper-
ature and humidity ratio to decrease simultaneously. The exhaust air can reach temperatures
lower than the freezing point of water, since frosting only starts when the air is saturated with
water vapour. It can be seen from Figure 3.12 and Figure 3.13 that the key ingredients for
frosting are a high humidity ratio of the exhaust air and low surfaces temperatures inside the
heat exchanger.
The outdoor temperature above which frosting does not occur is called the frosting limit
and is dependent of the type of heat exchanger. For heat exchangers with sensible heat transfer
only, this frosting limit is typically around 5 C [43]. The formation of frost in heat exchangers
with moisture transfer, for example energy wheels, is less straightforward. A study by Bilodeau
et al. [44] investigated frosting in energy wheels. The time until frost is formed in the energy
wheel is shown in Figure 3.14 for different outdoor temperatures and humidities. It can be seen
that, similar as with a plate heat exchanger, frosting occurs especially with high indoor relative
CHAPTER 3. ENERGY RECOVERY VENTILATION 18

Figure 3.12: Frost formation in a heat exchanger without moisture transfer [42]

Figure 3.13: Frost formation in a heat exchanger with moisture transfer [42]
CHAPTER 3. ENERGY RECOVERY VENTILATION 19

humidities and low outdoor temperatures.

Figure 3.14: Duration to frost formation in an energy wheel for different outdoor temperatures and
humidities [44]

Two frost-protection techniques and defrosting methods have be evaluated by Rafati Nasr
et al. [45]. The first method is preheating the supply air when the outdoor air temperature
falls below the frosting limit. This preheating by electricity or gas requires auxiliary energy,
but has the advantage that the heat exchanger operation is continuous. The other option is to
bypass the supply air to an auxiliary heater. The exhaust air then melts the frost in the heat
exchanger. After this defrosting period, the bypassing is turned off and the normal operation is
continued. Other methods, described by Kassai et al. [40], are stopping the air supply (building is
depressurised) or heating the core of the heat exchanger. Finally some more possibilities include
the reduction of the supply airflow rate, recirculation of the indoor air (no air leaves or enters
the building through the ventilation system), reducing the effectiveness of the heat exchanger
(changing the rotational speed for an energy wheel) or use a double core heat exchangers where
both cores are alternately used [42].
Chapter 4

Human influences

The accuracy with which temperatures, air flows and energy demand can be predicted, does not
only depend on the quality of model of the building and HVAC-system. The behaviour of the
occupants has a great influence on the energy demand [46].

4.1 User behaviour


To improve the accuracy of energy simulations, a correct understanding of user behaviour is
necessary. The effects of occupant behaviour on energy use were investigated by Van Raaij
and Verhallen [47]. A survey about the use of the heating system, the use of the mechanical
ventilation system, the number and duration of showers and the use of appliances was given
to the respondents. Based on the answers, behavioural patterns and user profiles were defined.
These behavioural patterns have the objective of categorising groups of related behaviour. The
five defined behavioural patterns, based on use of ventilation and heating system, are:

Conservers: People with a conscious use of energy.

Spenders: High temperatures and level of ventilation. Often people in houses with bad
insulation or people that are home often.

Cool : High ventilation and low indoor temperatures, resulting in an intermediate energy
use.

Warm: Typically older people with a high emphasis on comfort. Approximately the same
energy use as with cool.

Average: Intermediate temperatures and level of ventilation. The energy use is typically
higher than with cool or warm.

How people in these patterns handle heating and ventilation, is visualised in Figure 4.1. Similar
behaviour patterns were later used by Paauw et al. [48] and Santin [49]. The author also defined
groups of households with similar characteristics that behave in a similar fashion, called user
profiles. The four user profiles are single or low income couple, high income couple, family and
seniors. Energy use was compared for the different behavioural patterns and user profiles. As
seen in Figure 4.2, no significant differences between energy use for the behavioural patterns
were found.
Energy demand does depend on the user profiles. Singles and low-income couples are the
user profile that use least energy. This is shown in Figure 4.3. High-income couples tend to
have a more energy intensive behaviour, families use more appliances and seniors have higher
temperatures for a longer time. This causes them to have a higher energy usage.

20
CHAPTER 4. HUMAN INFLUENCES 21

Figure 4.1: Use of heating and ventilation for the different behavioural patterns [47]

Figure 4.2: Energy used for heating for different behavioural patterns [49]

Figure 4.3: Annual energy usage for different user profiles [49]
CHAPTER 4. HUMAN INFLUENCES 22

In a study in Austria by Haas et al. [50], the impact consumer behaviour and the thermal
quality of the house (indication of the level of insulation and airtightness, expressed in W/m2 K)
on residential energy demand for space heating is investigated. Since, with increasing thermal
quality, the heat load (for heating to the same temperature) should decrease, a linear relationship
between thermal quality and energy demand is expected, but not observed. For people living
in well-insulated houses, the price to heat to higher temperatures is lower and a higher comfort
level is demanded. The observed increase in energy demand is called the rebound-effect. The
rebound-effect is also observed in other studies [51; 52]. According to Haas and Biermayr [52],
this rebound-effect could amount to 20-30%. The rebound-effect can be seen in Figure 4.4 (SFD
= Single-family dwelling, MFD = Multi-family dwelling).

Figure 4.4: Visual representation of the rebound-effect [50]

The relationship between choice of indoor temperature on the energy demand was also inves-
tigated. A linear correlation, represented in Figure 4.5, was found. The author claims that the
choice of indoor temperature is the prevailing feature of consumer behaviour and vice versa. As
seen in Figure 4.6, the choice of indoor temperature increases as the building quality increases.

Figure 4.5: The relationship between the choice of indoor temperature and energy demand [50]

To do energy simulations, occupancy patterns need to be generated for each of the studied
user profiles. This modelling is usually done based on data from time-use surveys, as for example
by Aerts et al. [53] in Belgium. To asses internal gains, electricity load profiles are required. For
CHAPTER 4. HUMAN INFLUENCES 23

Figure 4.6: The relationship between thermal quality of the building and the choice of indoor temper-
ature [50]

the development of these profiles, time-use surveys and statistical data on electricity demand
and common households appliances are used [48; 54; 55]. A comparison of the electricity load
profile from the model of Yao and Steemers [55] with the load profile created from statistical
data is given in figure Figure 4.7. It shows low loads when people are usually sleeping and the
highest loads in the evening when people are performing activities like cooking and watching
television.

Figure 4.7: Comparison of two load profiles [55]

Next to occupancy patterns and electricity load profiles, heating profiles are required. Del-
ghust et al. [56] compared occupancy patterns and heating profiles. The authors conducted two
case studies in Belgium.
The first case study investigated 36 houses in an old social housing neighbourhood. Some
properties of these houses are bad thermal insulation, naturally ventilated and usually only heat-
ing in the living room (and decentralised electric heating in some other rooms). The occupancy
and heating profiles are shown in Figures 4.8 and 4.9. A strong correlation is observed between
presence and heating for the living room. This correlation is even stronger when comparing the
heating of the living room with the probability of having at least one person present in the
the living room. There is also a correlation between presence and heating for the bathroom.
CHAPTER 4. HUMAN INFLUENCES 24

Figure 4.8: Probability of presence for the first case study [56]

Figure 4.9: Probability of heating for the first case study [56]
CHAPTER 4. HUMAN INFLUENCES 25

The electric heaters present in the bathrooms of this neighbourhood are turned on, only when
that bathroom is used. Delghust et al. [56] also compared the daily number of heating hours
of different rooms for both neighbourhoods. The duration of heating hours for the living room
is compared with the duration of heating hours for the kitchen and bedroom in Figure 4.10.
The size of the dots represent the number of case studies situated on this point. The absence
of a heating system in the kitchen explains the horizontal line in Figure 4.10 (left). There is a
possibility for heating the bedrooms, but as the horizontal line in Figure 4.10 (right) indicates,
most households do not heat their bedrooms.

Figure 4.10: Comparison of the heating duration of the living room and kitchen (left) and the bedroom
(right) for the first case study [56]

The second case study is a recently built neighbourhood of 26 privately owned houses. These
houses have good insulation, mechanical ventilation and central thermostat heating system. The
occupancy and heating profiles of this neighbourhood are shown in Figures 4.11 and 4.12.

Figure 4.11: Probability of presence for the second case study [56]

For this neighbourhood, the correlation between presence and heating for the living room is
comparable to the first neighbourhood. The probability for heating of the kitchen and bathroom
is higher than the old neighbourhood. This probabilities are not depending on presence in the
CHAPTER 4. HUMAN INFLUENCES 26

Figure 4.12: Probability of heating for the second case study [56]

kitchen or bathroom. This is explained by the synchronous heating by the central thermostat.
The comparison of heating hours for the different rooms of this case study is shown in Figure 4.13.

Figure 4.13: Comparison of the heating duration of the living room and the kitchen (left) and the
bedroom (right) for the second case study [56]

The fact that almost all households heat their kitchen together with their living room can
clearly be seen in Figure 4.13 (left). Most of the the master bedrooms are not heated, which is
similar as for the old neighbourhood.
CHAPTER 4. HUMAN INFLUENCES 27

4.2 Thermal comfort


People change thermostat and ventilation settings to feel comfortable in the house. Various
studies investigated the quantification of comfort

4.2.1 Standard approach


Fanger et al. [57] developed a heat balance thermal comfort model. According to this model,
obtained from climate chamber experiments, comfortable skin temperature and sweat secretion
lie within narrow limits. Expressions for thermal comfort for different metabolic rates, clothing
insulation and environmental conditions were proposed. Metabolic rate, the heat produced by
a human body, depends on the performed activity. It usually expressed in mets, where the
standard value for a person at rest corresponds to one met (approximately 105 W ). The level of
clothing insulation is expressed by the unit clo (approximately 0.115 m2 K/W ). Typical values
are given in Table 4.1 and Table 4.2, respectively.

Table 4.1: Metabolic rates for different activities [58]

Activity Activity level (mets)


Reclining 0.8
Rest or seated 1.0
Light office work 1.1
Moderate office work 1.25
Light bench work 2.1
Moderate bench work 2.8

Table 4.2: Thermal resistance of clothing [58]

Clothing level R-value (clo)


Summer dress 0.35 - 0.45
Lightweight office dress 0.55 - 0.7
Medium-weight office dress 0.7 - 0.9
Business suit 0.9 - 1.3

Using Fangers expressions, the ASHRAE created a 7-point system between 3 (feeling
cold) and +3 (feeling hot) [59]. The neutral sensation is indicated by 0. The comfort vote for
a large group of individuals with the same clothing insulation and the same metabolic rate for
a given environmental condition is called the Predicted Mean Vote (PMV) of this condition.
A more meaningful quantitative measure is the Predicted Percentage Dissatisfied (PPD), since
this indicates what fraction of a group of people would dislike being in a certain condition.
The relationship between PMV and PPD was determined by Fanger et al. [57] and is shown in
Figure 4.14. A first observation on Figure 4.14 is that a condition where everybody is satisfied
(P P D = 0) does not exist. Next, it should be noted that the curve is symmetric around the
PMV-value 0, so in neutral conditions the amount of people that feel hot will be the same as
the amount of people that feel cold.
Using PMV and PPD, the International Standard ISO 7730 [61] predicts the thermal sen-
sation of people in moderate conditions and specifies comfortable conditions. Although the
standard takes local discomfort and draught into account, different studies have indicated flaws
in the standard [62; 63].
CHAPTER 4. HUMAN INFLUENCES 28

Figure 4.14: PPD in function of PMV [60]

4.2.2 Alternative approach


To cope with imperfections of the previously described methods to asses thermal comfort, adap-
tations have been proposed. According to Humphreys and Fergus Nicol [62], predictions based
on the PMV-value are biased with respect to operative temperature, humidity, air movement,
clothing insulation, metabolic rate and outdoor temperature. This bias is caused by the fact that
Fangers experiments were steady-state laboratory experiments with people wearing standard-
ised uniforms and executing sedentary activities. The authors state that accuracy of PMV can be
limited and thus, since PPD is affected by PMV, misleading when predicting the dissatisfaction
of people in everyday conditions.
Brager and de Dear [63] commented that PMV fails because adaptation was not taken into
account. People create their own thermal preferences by interacting with the environment, modi-
fying their own behaviour and gradually altering their expectations to adapt to the environment.
This thermal adaptation can be attributed to three different processes. Behavioural adjustment
of the bodys heat balance consists of the modifications a person makes in a certain setting.
Some examples are adjusting clothing, opening a window and scheduling activities. These mod-
ifications might be made both consciously or unconsciously. Next, physiological acclimatization
is defined as all changes in the physiological responses to gradually diminish the strain induced
by exposure to a thermal environment. This is possible with time scales beyond the life of an
individual (genetic adaptation) or over a period of days to weeks (acclimatization). These adap-
tations are changes in the nervous system and directly affect physiological thermoregulation set
points. The effects of cognitive and cultural variables are finally called psychological feedback.
The history of past and current thermal experiences with an indoor and outdoor climate can
affect the thermal sensation and feeling of discomfort.
To study comfort in residential buildings, Peeters et al. [60] divided houses in three thermal
zones: bathrooms, bedroom and other rooms. The authors took adaptation to the outdoor
climate into account by defining a reference external temperature Te,ref ( C):

Ttoday + 0.8 Ttoday1 + 0.4 Ttoday2 + 0.2 Ttoday3


Te,ref = (4.1)
2.4
In this formula, Ttodayi is the arithmetic average of the maximum and minimum external
temperature of i days ago.
When describing comfort in the bathroom, a critical lower limit was defined as the coldest
temperature acceptable to a nude, wet body, however still comfortable for a dried and clothed
person. Based on own research and results of other studies [6466], the authors derived equations
CHAPTER 4. HUMAN INFLUENCES 29

for the neutral comfort temperature Tn ( C):

Tn = 0.112 Te,ref + 22.65 C for Te,ref < 11 C (4.2)



Tn = 0.306 Te,ref + 20.32 C for Te,ref 11 C (4.3)

Occupants of bedrooms have a different metabolic rate and clothing value than people in an
office environment. The Fanger equations should thus not be used to asses comfort in bedrooms.
The warm temperatures in bedrooms in summer have a strong effect on sleep quality. CIBSE
[67] reports a steep drop in sleep quality when the temperature in the bedroom rises above
24 C and suggest a maximum temperature of 26 C. It should be noted, that if a fan is used,
an increase up to 3 C is acceptable. During winter two types of bedrooms are distinguished.
Bedrooms also used for activities as watching TV or doing homework are heated to temperatures
comparable to living room temperatures, measured in the same period. Bedrooms only used for
sleeping are often non-heated, resulting in cold conditions. A comfort temperature of 17 C is
indicated by CIBSE. A lower limit is set by the WHO recommendation of 16 C, because of a
lower resistance to respiratory infections for lower temperatures. Neutral comfort temperatures
have again been derived by Peeters et al. (for bedrooms with no fan in summer):

Tn = 16 C for Te,ref < 0 C (4.4)



Tn = 0.23 Te,ref + 16 C for 0 C Te,ref < 12.6 C (4.5)

Tn = 0.77 Te,ref + 9.18 C for 12.6 C Te,ref < 21.8 C (4.6)

Tn = 26 C for Te,ref 21.8 C (4.7)

Conditions in other rooms, like the living room and the kitchen, are comparable to office sit-
uations. Metabolic rates are usually slightly higher than in office environments and there is a
higher dependency on the outdoor climate, since people have more options to adapt. Some adap-
tations that can be made in residential houses are changing activity or clothes, switching rooms,
opening doors or windows and drinking hot or cold beverages. Because of these differences, and
the observation that, at the same temperature, people tend to feel warmer in their own house
(possibly by the presence of furnishings), Peeters et al. derived following neutral temperatures:

Tn = 0.06 Te,ref + 20.4 C for Te,ref < 12.5 C (4.8)



Tn = 0.36 Te,ref + 16.63 C for Te,ref 12.5 C (4.9)

With equations discussed in this section and the lower and upper limits for 10% and 20% P P D
given in the same study by Peeters et al. [60], comfort in all rooms of a building can assessed
for different outdoor temperatures.
Part II

Model description

30
Chapter 5

Building

To asses the performance of energy recovery ventilation, three buildings are studied: a detached,
a semi-detached and a terraced building. The current chapter describes these three buildings,
while in Chapter 6 and Chapter 7 the HVAC-system and the occupancy of the buildings are
discussed. As will be seen throughout the current chapter, the influence of several parameters on
the recovery performance is examined. These parameters are: the type of building, the nominal
energy demand for heating, the ventilation flow rates, the ventilation strategy, the effectiveness
of the heat exchanger, the occupancy of the building and the demanded comfort level. These
are explained throughout the model description and summarised in Chapter 8. There is one
base case defined with these parameters. This base case will be used the evaluated the influence
of each parameter. How this base case is defined is explained in the following chapters.
This study uses TRNSYS to simulate the buildings described in the next chapter and to
calculate the heat demands. TRNSYS is a simulation software package for transient systems
[68]. Results are needed for one year, but the simulations are done over one year and one week
to create a balance in the house after this first week. The time-step of the simulations is 15
minutes. A complete description of the TRNSYS model is given in Appendix D.

5.1 Building types


The investigation of the impact of the building type on the heat recovery performance is done
by comparing three different types of houses: a detached house, a semi-detached house and a
terraced house. The detached house will be used as base case, since this type of buildings is
most occurring in Belgium [69]. In Table 5.1, the gross and net floor area Af loor,g and Af loor,n ,
the interior volume (V ), the external wall area (Awall ) and the roof area (Aroof ) are given.

Table 5.1: Building properties

Af loor,g Af loor,n V Awall Aroof


Building type m2 m2 m3 m2 m2
Detached 176.4 137.0 378.9 187.8 88.6
Semi-detached 187.8 147.9 407.4 116.7 87.8
Terraced 174.5 126.3 357.1 83.4 51.7

The three houses have the same number of rooms. Each house has a living room, a study, a
toilet, a kitchen, a storage room, three bedrooms, a bathroom and a hall. Each building has an
attic as well, but this is not included in the conditioned volume.
The detached house is a two-storey building without adjacent neighbours. The semi-detached
house, also two storeys, has one adjacent neighbour, located in the North. The terraced house
shares a wall with its East and West neighbours and has three storeys.

31
CHAPTER 5. BUILDING 32

A simplified floor plan of the detached building is given in Figure 5.1. On this figure, the
North is located at the bottom of the plan. More detailed floor plans of all three buildings are
given in Appendix A.

(a) Ground floor (b) First floor

Figure 5.1: Floor plan of the detached building

5.2 Energy demand for heating


Three versions of each type of building are created, differing in the annual energy demand for
heating and created by varying the level of insulation and airtightness. The version with an en-
ergy demand of 15 kW h/m2 a (kW h/m2 per year) represents a passive house [70]. 30 kW h/m2 a
corresponds to a low-energy building [70]. Lastly, a building with an energy demand for heating
of 60 kW h/m2 a is modelled. This last version corresponds most to the current building stock
[71] and will therefore be used in the base case. To make these models, first the airtightness is
changed. Then the insulation is iteratively adapted until the building has the desired energy
demand for heating under the set of conditions described next.

5.2.1 Test conditions


The conditions used to create the three versions of the three building types are the same for each
building. All rooms are continuously heated to their set-point temperatures. The temperatures
to which each room is heated, are given in Table 5.2.
Table 5.2: Set temperatures for determining heat demand

Room Temperature ( C)
Living 20
Study 20
Kitchen 20
Toilet 18
Hall 18
Bedrooms 20
Bathroom 24

The internal heat gains and the gains by presence of people used for the construction of the
models are the same as the ones used with the simulations. The gains and occupancy patterns
are described further (Chapter 7).
Although the focus of this thesis is on buildings with both mechanical supply and extraction
(type D ventilation system), also buildings with natural supply are studied (type C ventilation
CHAPTER 5. BUILDING 33

system). The insulation thickness is determined with the type D buildings and the same insula-
tion is the used for the corresponding buildings with a type C ventilation system. The buildings
with mechanical supply and extraction are equipped with a heat exchanger, which should have
the same effectiveness for the three building types. The effectiveness is however not the same
for the three energy demands. The sensible effectivenesses used are 62.5%, 75% and 84%, for
the 60kW h/m2 a, 30kW h/m2 a and 15kW h/m2 a versions of the houses respectively.

5.2.2 Insulation
To achieve these levels of energy demand, the thickness of insulation in the walls, the roof and the
floor was varied. This was done iteratively until the heat demand approximated 15 kW h/m2 a,
30 kW h/m2 a and 60 kW h/m2 a for the three buildings. The resulting thickness and heat transfer
coefficients of the different parts of the building are given in Appendix B.

5.2.3 Infiltration, exfiltration and internal air flows


To account for the air flows through cracks in the building envelope and between the rooms, a
network of cracks is added, as is typically done with TRNSYS [68]. Air flow through walls and
the roof is modelled with two cracks to allow for the stack effect (rising of warmer and lighter
air). One crack is placed at 1/4 of the wall height, while the second crack is located at 3/4 of
the wall height. For modelling the air flow through the ceiling, one crack is used [72].
The leakage characteristic of cracks is described by a power law:

m = Cs (p)n (5.1)

In this equation, m is the air mass flow rate through the crack in kg/s, Cs is the air mass flow
coefficient (kg/s @ p = 1P a), p is the pressure difference over the crack (P a) and n is the
air flow exponent (-), which is approximately 0.5 for turbulent flow and 1 for laminar flow. In
this model, a value of 0.66 is used, since most cracks have a mixed flow regime which is situated
between 0.6 and 0.7 [68].
The air mass flow coefficient determines the flow rate through a crack for a certain pres-
sure difference. By manipulating this factor, the airtightness (= n50 , as described earlier in
Section 3.2.1) of the house can be varied. The same airtightness is used for the three building
types, but it is changed for each version of the house. The houses with an energy demand of
15 kW h/m2 a, 30 kW h/m2 a and 60 kW h/m2 a are modelled with an airtightness of respectively
0.6 ach, 3 ach and 6 ach, for a pressure difference of 50 P a between inside and outside the build-
ing. How this Cs -value is calculated, is explained in Appendix C.

The Belgian standard NBN D 50-001 [15] specifies that buildings should be equipped with
flow openings that act as natural exhaust for the dry rooms (e.g. living and bedrooms) and as
natural supply for the wet rooms (e.g. bathroom and kitchen). These flow openings must allow
an air flow of 25 m3 /h at a pressure difference of 2 P a, except for the kitchen, where the air
flow should be 50 m3 /h at the same pressure difference. Since these flow openings have a more
turbulent flow than a crack in a wall, a flow exponent, of 0.5 is chosen. The resulting air flow
coefficient, Cs , is 5.893 103 kg/m2 s @ 1 P a.
Chapter 6

HVAC-system

This chapter first deals with the heating and ventilation systems. To investigate energy recovery
performance, the buildings are equipped with system D ventilation. A comparison is made with
another type of ventilation system, with mechanical extraction but natural supply (system C).
Other parameters related to the ventilation system that are discussed are the ventilation flow
rate, the heat exchanger effectiveness and the effects for demand-controlled ventilation. The
chapter ends with a discussion of the heating strategy.

6.1 Ventilation
6.1.1 Flow rates
For the base case, the ventilation flow rates prescribed by the standard NBN D 50-001 [15] will
be used. To investigate the effects of using the ventilation system on medium or low setting,
simulations using 2/3 and 1/3 of the nominal flow rate (high setting) will be done.
The nominal flow rates, required by the standard for the different buildings and ventilation
systems are given in Table 6.1. The required supply air flow rates are the same for system C
and D. Since ventilation needs to be balanced with system D, the exhaust flow rates are higher
for system D than for system C.

Table 6.1: Mechanical ventilation flow rates used in the buildings as prescribed by standard NBN D
50-001 [15]

Detached Semi-detached Terraced


Exhaust Supply Exhaust Supply Exhaust Supply
Room C D C D C D
Living room 132 102 119
Study 30 62 54
Kitchen 50 81 50 80 50 80
Storage room 50 75 50 75 50 75
Toilet 25 35 25 35 25 35
Hall 0 70 0 70 0 70
Bedroom 1 63 63 65
Bedroom 2 69 72 47
Bedroom 3 67 51 72
Bathroom 50 100 50 90 50 97
Total 175 361 361 175 350 350 175 357 357

34
CHAPTER 6. HVAC-SYSTEM 35

6.1.2 Flow controllers


For buildings with a type D ventilation system a flow controller is not required, but it is necessary
for the versions of the buildings with system C ventilation, because there, no fan is present to
supply fresh air to the living room, study and bedrooms. A flow controller is fully open when
there is no pressure difference over it and gradually closes with increasing pressure difference.
These flow controllers are demanded by the standard NBN D 50-001 [15], which says that at a
pressure difference of 2 P a over the controller, at least the flow rates listed in Table 6.1 should
reached. The flow rate (in m3 /h) is modelled with an exponential characteristic, similar to that
of cracks: Q = Cq (p)n , with Cq the air volume flow coefficient (m3 /h @ p = 1 P a) and
n = 0.66 [68]. With the flow rates from Table 6.1, the Cq -value for the rooms can be calculated
as: Cq = Qnom /20.66 , with Qnom the nominal flow rates from Table 6.1. The resulting values
are listed in Table 6.2.
Table 6.2: Flow coefficients for the flow controllers in the buildings with system C ventilation, according
to standard NBN D 50-001 [15]

Cq , (m3 /h at 1 P a)
Detached Semi-detached Terraced
Living room 83.54 64.55 75.31
Study 18.99 39.24 34.18
Bedroom 1 39.87 39.87 41.14
Bedroom 2 43.67 45.57 29.75
Bedroom 3 42.40 32.28 45.57

6.1.3 Heat exchanger


For this study, a sensible-only heat exchanger is used in the buildings equipped with a type D
ventilation system. With a sensible-only heat exchanger, no moisture transfer is possible from
one air stream to the other. A sensible effectiveness of 75% at nominal flow rate is chosen, based
on the EPB-database [73].
To investigate the sensitivity of the sensible effectiveness on the energy demand for heating,
simulations are also done with a sensible effectiveness of 80% and 85%.
There is also an influence of the ventilation flow rate on the effectiveness. When the air flow
rate decreases, the effectiveness of the heat exchanger increases. The efficiency is modelled with
the -NTU correlations for a crossflow heat exchanger [32]. These correlations are given by
equations (6.1) and (6.2).
1exp[N T U (1C )]
= 1C exp[N T U (1C )] , C 6= 1 (6.1)
= 1+N NT U
TU , C = 1 (6.2)
With balanced ventilation, the supply and exhaust flow rates are approximately the same, which
A is calculated based on the flow rates
means C = 1. The number of transfer units, N T U = CUmin
and the effectiveness in design conditions:
d
N T Ud = (6.3)
1 d
U A = (U A)d = N T Ud Cd (6.4)
(U A)d
NTU = (6.5)
Cmin
Now N T U is calculated with (6.5) and used in (6.1) or (6.2) to find the effectiveness with the
new air flow rates. From these equations, it can be derived that when the flow rate is reduced to
CHAPTER 6. HVAC-SYSTEM 36

2/3 or 1/3 of the nominal flow rate, the effectiveness increases to 81.8% and 90.0% for a design
effectiveness of 75%.

6.1.4 Demand-controlled ventilation


In Section 2.4, demand-controlled ventilation (DCV) was already discussed. CO2 -based DCV is
a possibility with the model, but is not used in the base case. In the base case, ventilation is
always on, regardless of the CO2 -concentration.
As atmospheric concentration of CO2 , 400 ppm is chosen. The production rate by people
is set to 1.2 105 kg/s per active person and is lowered to 6.7 106 kg/s for a sleeping person
[74]. The ventilation strategy used here is a simple on/off-strategy. When the maximum CO2 -
concentration in the house rises above 1000 ppm, ventilation is turned on to the desired flow
rate. When this concentration drops below 900 ppm all ventilation flow rates are reduced to
10% of the nominal air flow rate.

6.2 Heating
For this study, a heating strategy is chosen where not all rooms are heated simultaneously. A
room is instantly heated (unlimited heating power) to the desired temperature only when people
are present in this room (occupancy of the building is discussed in Chapter 7). By the unlimited
heating power, the heat demand might be different than without unlimited heating power, but
the desired comfort temperatures are always reached. Since the occupancy is divided over the
living room, study, kitchen, one bedroom and the bathroom, only these five rooms are heated.
The rest, being the storage room, toilet, hall, other bedrooms and attic, are non-heated rooms
The temperatures to which these rooms are heated is dependent on the desired comfort level
and are based on the study by Peeters et al. [60], as described in Section 4.2.2. In this study,
an outdoor reference temperature is calculated, based the outdoor temperature of the current
and previous days. Correlations are given for a neutral temperature in the bathroom, bedroom
and other rooms in function of this reference temperature, together with 10 P P D and 20 P P D
(Predicted Percentage Dissatisfied) temperature ranges. Three comfort levels will be compared.
The operative temperature is used to asses comfort and is based on the air temperature and
the mean radiant temperature. It is used because only the air temperature is not sufficient
to determine comfort. Warm surfaces can cause a person to feel warmer even though the
surrounding air is cold, and cold walls can create a feeling colder than the air temperature. This
is expressed by the mean radiant temperature [75].
For the base case, the rooms are heated so that the operative temperatures are the same
as the neutral temperatures. For the two other desired comfort levels, the rooms are heated to
either the lower or upper bound of the 10 P P D temperature range.
Chapter 7

Internal heat gains and occupancy

7.1 Internal heat gains


In the house, several sources of heat are present that contribute to the heating of the building.
The sources of internal heat gains used in this model and the heat they produce are given in
Table 7.1. These values are based on a study by de Meester et al. [76]. The total values are

Table 7.1: Sources and power of the internal heat gain

Source Power
Fridge 0.85 kW h/day
Washing-up 0.33 kW h/use, 65 uses/(year person)
Appliances 50 kW h/(year person)
Television 150 W
Computer 70 W
Cooking 912 W
Lighting 6 W/m2
Shower 1486 W/shower

divided with a time-step of 15 minutes over the living room, kitchen, study, bathroom and
bedroom with a one-day schedule, which is repeated for the whole year. For one year, these heat
gains of appliances amount to 1582.74 kW h in total.
Also human bodies produce heat. This is modelled by 75 W when people are awake and 70 W
for a sleeping person [58]. With the occupancy pattern of the base case (occupancy patterns are
explained in the next section), a total amount of 854.85 kW h of heat is released by the presence
of people.
These internal gains would have a greater impact on the smallest building. Therefore the
heat gains by appliances are scaled so that the sum of internal heat gains and gains by people
are on average 2.1 W/m2 (net surface area) for each house. This is the value used by the Passive
House Planning Package (PHPP) [77].

7.2 Occupancy
As explained in Section 4.1, user behaviour can influence the heat demand of a building. This is
modelled with two occupancy profiles. The profiles are based on the study by Aerts et al. [53].
With the data from a time-use survey in Belgium, seven occupancy patterns with three possible
states were created, and are shown in Figure 7.1. The bottom curve indicates the probability
of someone being at home and awake at a certain moment. The vertical distance between the
first and the second curve is the probability of someone being at home but asleep, while the

37
CHAPTER 7. INTERNAL HEAT GAINS AND OCCUPANCY 38

probability of no one being at home is indicated by the distance between the second curve and
1.

pattern 1: pattern 2: pattern 3: very short pattern 4:


mostly absent mostly at home daytime absences night-time absence
1 1 1 1

0.5 0.5 0.5 0.5

0 0 0 0
4AM 12PM 20PM 4AM 4AM 12PM 20PM 4AM 4AM 12PM 20PM 4AM 4AM 12PM 20PM 4AM

pattern 5: pattern 6: pattern 7:


daytime absence afternoon absence short daytime absence
1 1 1

0.5 0.5 0.5

0 0 0
4AM 12PM 20PM 4AM 4AM 12PM 20PM 4AM 4AM 12PM 20PM 4AM

Figure
Figure 7.1: 5: Average
Seven occupancy
occupancy patterns
patterns from
defined by clustering
Aerts et al. [53]
between 25 and 39 years old (n=1000). On the left, the average occupancy pattern is shown.
In general, the correspondence between the observed (coloured) and simulated (dotted) data is
Insatisfactory.
this study,Fortwotheopposite patterns,
simulations of this pattern 5 and pattern
type of household, 2 are
we first used to the
determined model the occupancy
distribution of
clusters based
on weekdays. on base
In the the individuals age (table 1).consists
case, the household The deviations between observed
of two members, which and
are simulated
mostly absent
during data can be explained
daytime on weekdays by the fact that5).
(pattern we only
When used
theage to assign aparameter
occupancy cluster to each simulated
is varied, the two
individual.
occupants The results
are mostly at home may improve
during when increasing
weekdays (pattern 2).the number
In both of parameters.
cases, weekendsTheareright-
modelled
with hand side of figure pattern,
an intermediate 6 illustrates the average
pattern results
3, where for two
there are activities: preparingAfter
short absences. food (task) and the
one week,
taking pattern
occupancy a bath/shower (personal activity). The red and green lines correspond to the observed and
is repeated.
simulated
Since data,for
the data respectively.
the profiles Wefrom
notice
thestrong
studyagreement
by Aertsbetween the isobserved
et al. [53] and simulated
given with a time step of
data. The simulated data appears to be more smoothed out than
10 minutes and the simulations in this study are done with a time step of 15 minutes, the observed data. On the the
one data
is firsthand, this can betoexplained
transformed this newbytimethe fact
step.that the original dataset selection (2-person households
with members between 25 and
Because it is not possible that at a certain 39) only consists of 398e.g.
instant, households, whereas
0.37 people areweawake
simulated 1000house,
in the
the resolution of the data is adapted. Since there are five weekdays and the household off,
synthetic households. On the other hand, the start- and duration probabilities were rounded has two
since they were calculated for each half hour instead of each (10-minute) time
members, the highest possible resolution is 1/(5 2) = 0.1, so all probabilities are rounded up to step.
one decimal.
Average occupancy Preparing food Taking a bath or shower
Finally, the occupancy is manually divided over the five weekdays and five rooms (living room,
1 0.2
kitchen, study, observed data
absentbedroom and bathroom). For example, if the probability for someone being at
0.1
0.15
home and awake is 1 at a certain time, there would be two people awake in the house from
sleeping
simulated data
Monday0.5 to Friday at this moment 0.1 of the day. If the probability were 0.3, this could be done
0.05
by one person being present from0.05Monday to Wednesday, but nobody on Thursday and Friday.
This method atishome used for patterns 2 and 5 (weekdays). Since a weekend only has two days, the
& awake
0 0
resolution
4AM of 12PM
the used 20PMdata4AM of pattern
4AM 3 is 0.25. 20PM
12PM 4AM 0
4AM 12PM 20PM 4AM
The original data of occupancy pattern 5, together with the data adapted for the new time step
Figure 6: Results
and resolution, is givenforintwo-person
Figure 7.2.households with both members between 25 and 39 years
In Figure 7.3 is shown howoccupancy
old. Left: the average much people pattern (coloured
are awake for =what
observed data,ofdotted
fraction the time linefor
= pattern
simulated data) Right: the average occurrence for the activities preparing
2, 3 and 5. When compared with Figure 7.1, it can be seen that the house in pattern 2 is indeed food and
taking a bath or shower (red = observed data, green = simulated data).
the most occupied, while the people of pattern 5 are the most absent. When nobody is awake,
everybody is either sleeping or out of the building. The households of pattern 2, 3 and 5 sleep
4 Conclusions
on average 8.4 hr, 9.9 hrand andFuture
8.0 hr, Work
respectively. Which rooms the people are in when they are
awake,Weisdeveloped
shown inaFigure 7.4. to obtain realistic individual household behaviour sequences for
methodology
building simulations. The method includes occupancy and activity sequences.
CHAPTER 7. INTERNAL HEAT GAINS AND OCCUPANCY 39

0.8

0.6

0.4

0.2
Original probability
Adapted probability
0
4AM 12PM 20PM 4AM

Figure 7.2: Original and adapted occupancy probabilities for pattern 5

Figure 7.3: How much people are awake what fraction of the time for the three occupancy patterns
used

Figure 7.4: Rooms occupied when people are at home and awake for the three occupancy patterns
Chapter 8

Parameter summary

The influence of eight parameters on the performance of energy recovery of the model described
above is to be investigated. Two of these parameters are related to the building, five to the
HVAC-system and the last one to the occupancy patterns. The parameters and the possibilities
for each parameter are listed here.

Building type

Detached building
Semi-detached building
Terraced building

Energy demand for heating

60 kW h/m2 a
30 kW h/m2 a
15 kW h/m2 a

Ventilation flow rate

Nominal flow rate (dictated by EPB-regulations)


2/3 of the nominal flow rate (medium setting of the ventilators)
1/3 of the nominal flow rate (low setting of the ventilators)

Ventilation strategy

Ventilation always on
CO2 -based demand-controlled ventilation

Sensible effectiveness of the heat exchanger

75%
80%
85%

Comfort level

Heat rooms to neutral temperature


Heat rooms to lower limit of 10 P P D range
Heat rooms to upper limit of 10 P P D range

40
CHAPTER 8. PARAMETER SUMMARY 41

Occupancy pattern

Daytime absence on weekdays (pattern 5)


Very short daytime absences on weekdays (pattern 2)

Ventilation system

Type D ventilation system


Type C ventilation system

Simulations are done with different combinations of these parameters. One base case is used
and referred to when other cases are studied. The base case is the detached building with a
nominal heat demand of 60 kW h/m2 a. It has a type D ventilation system using the nominal flow
rates continuously (no DCV). The people are mostly absent (pattern 5) and when at home, they
require a medium comfort level (heating to neutral temperatures). The sensible effectiveness of
the heat exchanger is set to 75%.
In the following chapters, the results are discussed.
Part III

Results and discussion

42
Chapter 9

Heat recovery performance

To evaluate the performance of the heat recovery system and compare different cases, several
efficiencies are calculated. These efficiencies are based on different definitions of the heat demand
of the building and the energy recovered by the heat exchanger. In the next section, the four
used definitions of the heat demand are explained: the actual heat demand, the heat demand
without a heat exchanger, the heat demand with a perfect heat exchanger and the heat demand
without ventilation. Also the definitions of the recovered energy and the maximally recoverable
energy are discussed. Finally, using these energy quantities, five different efficiencies are defined.

9.1 Heat demand


The actual heat demand, Q, is the amount of energy a building in a certain case uses for
heating the rooms to the desired temperature. It is the integral of the total heating power of
this building, P, over the heating season. How this heating season is determined is explained in
Section 9.2.

The heat demand without heat exchanger, Q0 , indicates how much energy a house would
require if no heat exchanger was installed. The same energy demand would be found if the
sensible effectiveness of the heat exchanger was reduced to 0%. This value will be higher than
Q and Q0 Q indicates how much energy is saved by using a heat exchanger.

The heat demand with a perfect heat exchanger, Q1 , is determined when the effectiveness
of the heat exchanger is increased to 100%. Since no energy is released in the atmosphere by
the ventilation system, Q1 will be lower than Q.

The heat demand without ventilation, Qnv , is obtained by turning off the supply and
exhaust fans. It will have the lowest value of the four heat demands, even though for both Q1
and Qnv , no energy is released in the atmosphere by the ventilation system. The difference
between Q1 and Qnv is that, with a perfect heat exchanger, the heat from the warmer rooms is
distributed to the non-heated rooms by the ventilation system and the colder air is distributed
to the heated rooms, which will require more energy to reach the desired temperature.

The calculation of the four heat demands is summarised in Figure 9.1, with the values of the
base case. In this figure, the integration of the heating power over the heating season is shown,
together with the final values. On figures showing the time on the x-axis, 0 hr corresponds to
1st January 00:00. The horizontal parts of the curves, e.g. the beginning and the end of Qnv
indicate that no energy is required for heating, where this was still the case for the building
under the same conditions, but without a heat exchanger (Q0 ).

43
CHAPTER 9. HEAT RECOVERY PERFORMANCE 44

50
Q
0
45
Q
Q1
40
Q nv

Heat demand (kWh/m)


35

30

25

20

15

10

0
-2000 -1000 0 1000 2000 3000 4000
Time (hr)

Figure 9.1: Determination of the four heat demands for the base case

9.2 Heating season


As can be seen in the description of the HVAC-system (Chapter 6), no cooling is present in
the buildings. In practice, the heat exchanger would be by-passed in summer to allow for free
cooling and reduce the risk of overheating. This study however investigates the performance of
heat recovery systems during the heating season, so the period outside this heating season, with
potentially overheating, is of no importance and is neglected.
A consequent definition of the heating season should be used to compare the results of the
different cases. For this study, this definition is based on the total heating power (sum of all
rooms) of the building from the studied case, but without the use of a heat exchanger (or a
sensible effectiveness of 0%). The heating season stops the first time of the year (starting on
1th of January) no heating is required for at least the next 24 hours. After the last time of the
year no heating is required for 24 hours or more, the heating season starts again. The heating
power of the base case without a heat exchanger is shown in Figure 9.2. The start and end of
the heating season, according to the previous definition, are indicated on this figure.

9.3 Recovered energy


Actual recovered energy
When the supply air is colder than the exhaust air, the cold air gets heated by the warmer
air in the heat exchanger. This amount of heat added to the supply air, is the energy actually
recovered by the heat exchanger, Qh , and it is the integral of the heat transfer rate, Ph , over the
heating season. The sensible effectiveness of the heat exchanger determines what fraction of the
exhausted energy is recovered. This recovered energy is supplied to the rooms by the ventilation
system and reduces the heat demand of the heated rooms.

Maximum recoverable energy


If a heat exchanger with a sensible effectiveness of 100% was installed, no energy would be
CHAPTER 9. HEAT RECOVERY PERFORMANCE 45

35

30 End of Start of
heating season heating season

Heating power, P0, (kW/m2)


25

20

15

10

0
0 1000 2000 3000 4000 5000 6000 7000 8000
Time (hr)

Figure 9.2: Heating power (in kW/m2 ), with start and end of the heating season indicated, for the base
case without heat exchanger

released in the atmosphere. The heat that in this case is added to the supply air, is the energy
maximally recoverable by the heat exchanger, Qhm , and equals the amount of energy that is ex-
tracted from the rooms by the ventilation system. Its value depends on the exhaust and supply
temperatures and the air flow rate. The higher the temperature difference between indoor and
outdoor and the higher the air flow rate, the greater this value would be.
In Figure 9.3, the integration of Ph and Phm to Qh and Qhm over the heating season is shown.

50
Qhm
Qh
40
Recovered energy (kWh/m2)

30

20

10

0
-2000 -1000 0 1000 2000 3000 4000
Time (hr)

Figure 9.3: Actual and maximum recovered energy integrated over the heating season
CHAPTER 9. HEAT RECOVERY PERFORMANCE 46

9.4 Efficiency definitions


With the four heat demands (Q, Q0 , Q1 and Qnv ), the recovered energy, Qh and the maximally
recoverable energy, Qnv , five different efficiencies are here defined. These efficiencies differ from
the sensible effectiveness of a heat exchanger in the sense that the effectiveness only considers
inlet and outlet conditions of the air in the heat exchanger at a certain instant, while the
efficiency looks at the different energy quantities over the complete heating season.
Firstly, the ratio of the usefully recovered energy, Q0 Q, to the maximal usefully recoverable
energy, Q0 Q1 , is called 1 .
Q0 Q
1 = (9.1)
Q0 Q1

The second efficiency, 2 is the ratio of the usefully recovered energy, Q0 Q, to the amount of
energy that is added to the supply air by the heat exchanger, Qh . This is the fraction of the
energy recovered by the heat exchanger that has usefully been supplied to the rooms.
Q0 Q
2 = (9.2)
Qh

Next, 3 is defined as the fraction of the heat extracted from the rooms, Qhm that has been
recovered in the heat exchanger, Qh .
Qh
3 = (9.3)
Qhm

The part that can be recuperated from the additional energy demand because of the use of
mechanical ventilation is indicated by 4 .
Q0 Q
4 = (9.4)
Q0 Qnv
The last efficiency that is defined, 5 , is the fraction of the energy in the air extracted from
the rooms, Qhm , that is first recovered by the heat exchanger and then usefully supplied to the
building, Q0 Q.
Q0 Q
5 = 2 3 = (9.5)
Qhm

In the following chapters, the performance of energy recovery ventilation will be evaluated
using these efficiencies. Firstly, the base case is discussed. Next, one parameter is varied, while
the others remain the same.
Chapter 10

Base case

In this study, the base case is defined as the detached building, with an energy demand for
heating of 60 kW h/m2 a (under the conditions described in Section 5.2.1), equipped with a type
D ventilation system. The ventilators are used on high setting and are always on (no DCV).
The heat exchanger has an effectiveness of 75%. The occupants are mostly absent and when at
home, they require a medium comfort level (heating up to neutral temperature, as defined in
Section 6.2).
How the different energy quantities and the different recovery efficiencies are determined is
explained in the previous chapter. The resulting heat demands and the recovered energy for
the base case are shown in Figure 10.1. The values of Qnv , Q1 , Q and Q0 are shown on the

50

40
Energy, kWh/m2

30
Q
nv
Q
20 1
Q
Q
0
10 Qh
Q hm

Figure 10.1: Heat demands, recovered energy and maximally recoverable energy for the base case

left graph, where it can be seen that for the base case, the building has a heat demand, Q, of
30.67 kW h/m2 . This value of Q is lower than the 60 kW h/m2 a to which the properties of the
insulation of the building were tuned, although the same building is used. This has different
causes. Firstly, only the living room, kitchen, study, bedroom and bathroom are heated for the
simulations, while every room was heated for the creation of the buildings. Next, also the heating
strategy has changed, since now a room is only heated when occupied instead of continuously.
The effectiveness of the heat exchanger is also increased from 62.5% for the construction of the
models (Section 5.2.1) to 75% for the simulations with the base case. Finally, the 30.67 kW h/m2

47
CHAPTER 10. BASE CASE 48

is only over the heating season and the 60 kW h/m2 a is for a whole year.
The division of the heat demand over the different heated rooms is given in Figure 10.2. The

Figure 10.2: Division of the heat demand over the heated rooms during the heating season for the base
case

high fraction of energy needed for the heating of the living room is explained by its large area and
by the high occupancy duration of the living room. This occupancy can be seen in Figure 7.4 on
page 39 in the part on the occupancy patterns. The low energy needed for heating the bedroom,
despite its high occupancy (awake and asleep), results from the lower required temperatures.
The temperatures to which the rooms are heated, is depending on the outdoor temperature,
but on average the set-point temperature in the bedroom is 18.0 C, while this temperature is
23.4 C and 21.0 C in the bathroom and other rooms respectively.
The values of Q0 and Q1 are 48.41 kW h/m2 and 24.91 kW h/m2 , so the heat demand is
decreased with 17.74 kW h/m2 or 36.65% by installing a heat exchanger which recovers the heat
of the exhausted air by bringing the supply air to higher temperatures. The heat demand
could even be decreased more with a better heat exchanger. A theoretical heat exchanger with
an effectiveness of 100% could lower the heat demand with an additional 5.76 kW h/m2 . The
reason Q0 is higher than Q and Q1 is lower, are the different temperatures of the air supplied
to the rooms. On average over the heating season, the temperatures of the fresh air leaving
the heat exchanger and supplied to the rooms for the base case with a heat exchanger with an
effectiveness of 75%, 0% and 100% are 15.2 C, 7.7 C and 18.2 C respectively.
The heat demand without ventilation, Qnv , is 16.83 kW h/m2 . As explained with the defini-
tion of the heat demands in Section 9.1, this is lower than Q1 , because of the redistribution of air.
With ventilation, the energy supplied to the heated rooms also elevates the temperatures in the
non-heated rooms, which is not the case when all ventilation is shut down. This is illustrated by
Figure 10.3 with the higher air temperature in the toilet (non-heated room) during the heating
season with ventilation. These higher temperatures in non-heated rooms cause more infiltration
and transmission losses.

The efficiencies of the ventilation system in the base case are shown in Figure 10.4. It can be
seen that 1 and 3 are both approximately equal to 75%, which is also the effectiveness of the
heat exchanger. A value for 3 near the sensible effectiveness strokes with the definition of S :
ratio of actual heat transfer to maximum possible heat transfer (see equation (3.1)). The value
CHAPTER 10. BASE CASE 49

28
Perfect heat exchanger
26 No ventilation

24

Temperature (C)
22

20

18

16

14

12

10
-2000 -1000 0 1000 2000 3000 4000
Time (hr)

Figure 10.3: Air temperature in the toilet during the heating season for the base case with a perfect
heat exchanger and without ventilation

of 1 implies that the heat demand reduction also scales with the effectiveness. This was tested
by simulating with different effectivenesses and the results are given in Figure 10.5. Both 1
and 3 vary linearly with the sensible effectiveness, S , of the the heat exchanger. The influence
of this effectiveness on the different energy definitions and the other efficiencies is explained
in Section 11.5. Although it was just seen that the useful recovered energy (Q0 Q) and the

100

80 75.49 75.31
Efficiency, %

60 56.17
49.87

40 37.56

20

0
21 22 23 24 25

Figure 10.4: Efficiencies for the base case

energy recovered by the heat exchanger (Qh ) scale in the same way, they are not equal to each
other. In this case, only 2 = 49.87% of the recovered energy contributes to the reduction of
the heat demand. Half of the recovered energy is supplied to rooms that do not require heating
at that instant (because it is either an non-heated room or a heated rooms which has already
reached its set-point temperature) and does, even with internal air flows, not reach rooms that
do. This energy first unnecessarily heats rooms and is then released in the atmosphere through
exfiltration, conduction through the walls or the ventilation system.
CHAPTER 10. BASE CASE 50

4 has a value of 56.17%. This means that the extra heat demand, required by mechanically
ventilating the house, Q0 Qnv , is reduced with 56.17% by using a heat exchanger.
The last efficiency to be discussed is 5 . Of all the energy that is extracted by the ventilation
system, only 37.56% is recovered and supplied to parts the building that need to be heated.

100 100
21 23

80 80
Efficiency 2, %

Efficiency 2, %
60 60

40 40

20 20

0 0
0 20 40 60 80 100 0 20 40 60 80 100
Sensible effectiveness 0 S, % Sensible effectiveness 0 S, %

Figure 10.5: Linear relationship between 1 and 3 and the sensible effectiveness of the heat exchanger,
S
Chapter 11

Influence of parameters on recovery


performance

The performance of the ventilation system with heat recovery in the base case was discussed in
the previous chapter. In this chapter, the influence of the different parameters is studied. Each
time, on parameter is changed with respect to the base case, while the rest remains the same.

11.1 Building type


The first parameter of which the influence on the performance of energy recovery is investigated,
is the building type. The base case, discussed in Chapter 10, studies a detached building, while
here a semi-detached and a terraced house are discussed. On Figure 11.1, the energy quantities
are given on the left and the efficiencies on the right for the base case, the semi-detached
building and the terraced building. It would be expected that the energy demand decreases for

60 100
Base case
Semi-detached building
50 Terraced building
80
Energy, kWh/m2

Efficiency, %

40
60
Q nv
30
Q1
40
20 Q
Q0
Qh 20
10
Q hm
0 0
BC SDB TB BC SDB TB 21 22 23 24 25

Figure 11.1: Heat demand and recovered energy (left) and efficiencies (right) for the base case (BC),
the semi-detached building (SDB) and the terraced building (TB)

the semi-detached building and even more for the terraced building, because these houses have
less external walls to the outside and more to an adjacent neighbour compared to a detached
house. However this trend is not observed on Figure 11.1. Instead of decreasing, the energy
demand for the semi-detached building has even increased with 1.16 kW h/m2 , compared to the
base case. The benefit of having lower transmission losses by a shared wall with an adjacent
neighbour was reduced in this study when the insulation thickness was tuned in Section 5.2.2,

51
CHAPTER 11. INFLUENCE OF PARAMETERS ON RECOVERY PERFORMANCE 52

to give the houses the same energy demand for heating under the conditions described there. It
can be seen in Appendix B, that the envelope of the detached building has the most insulation
(4.8 cm of polyurethane in the external walls), while the terraced building needs least insulation
(1.7 cm of polyurethane in the external walls) to achieve the 60 kW h/m2 a energy demand for
heating.
Also on Figure 11.1, it can be seen that Qh and Qhm are approximately the same for the base
case and the terraced house, but lower for the semi-detached house. This is explained by the
lower required ventilation flow rates for this last building (Table 6.1).
The efficiencies are shown on the right graph of Figure 11.1. As explained with the dis-
cussion of the base case, 1 and 3 are approximately the same and determined by the sensible
effectiveness of the heat exchanger. Compared to the detached building, the value of 2 is higher
for the semi-detached building but slightly lower for the terraced building. The lower flow rates,
causing the lower Qh , explain the increase of 2 for the semi-detached house. For the terraced
building, the reason for this slight decrease is the slightly lower value of Q0 -Q, with almost the
same value of Qh . The same trends can be seen with 5 as with 2 . The same explanation can
be used since 5 = 2 3 and 3 is only depending on the effectiveness of the heat exchanger.

11.2 Nominal energy demand for heating


To investigate the influence of the level of insulation and airtightness of the house, three versions
of the detached house are compared. The base case represents a house with an heat demand of
60 kW h/m2 a (under the conditions in Section 5.2.1). By increasing the level of insulation and
airtightness this value has been reduced to 30 kW h/m2 a and 15 kW h/m2 a.
The results for the three versions of the house is given in Figure 11.2. On the left graph, the
expected trend of a reduce in heat demand, Q, can be seen, from 30.7 kW h/m2 for the base to
18.7 kW h/m2 and 12.6 kW h/m2 for the two buildings with a lower nominal energy demand for
heating. The reason these values differ from 60 kW h/m2 a, 30 kW h/m2 a and 15 kW h/m2 a are
already explained with the discussion of the base case.
The same decreasing trend is present for the heat demand without ventilation, Qnv , also expected
when the insulation thickness is increased and the airtightness improved. The decrease of Q1
Qnv , which indicates the extra heat demand caused by the redistribution of the air over the
rooms, with increasing insulation, is attributed to the higher temperatures in the non-heated
rooms. For example, the average temperature of the hall over the heating season is 17.2 C for
the base case and rises to 18.6 C and 20.9 C for the two other cases, because of the better
insulation and airtightness.
An increasing trend is observed for Q Q1 and Q0 Q. Q0 Q is the usefully recovered
energy. This is higher when the insulation and airtightness improve, because less heat supplied
to non-heated rooms is lost through transmission and infiltration losses and more of this heat
contributes to the heat demand. Q Q1 represents the energy released in the atmosphere
because of the imperfection of the heat exchanger. Since the extracted air is warmer for a better
insulated house compared to the base case, more energy is discharged to the outdoors. This is
also expressed by Qh and Qhm in Figure 11.2. Higher indoor air temperatures, in the buildings
with more insulation and better airtightness, mean a higher temperature difference between
fresh and exhausted air, causing a higher heat transfer rate in the heat exchanger.

The performance of energy recovery for these buildings is shown on the right graph of Fig-
ure 11.2. Once again, 1 and 3 are equal to 75%, the effectiveness of the heat exchanger, for
the three cases.
The part of the recovery energy that is supplied usefully to the building, 2 , is approximately
50% for the three cases. Both Q0 Q as Qh increase, explaining this result. The same reasoning
is valid for the almost equal 5 .
CHAPTER 11. INFLUENCE OF PARAMETERS ON RECOVERY PERFORMANCE 53

70 100
Base case
60 30 kWh/m 2a
80 15 kWh/m 2a
50
Energy, kWh/m2

Efficiency, %
60
40
Q nv

30 Q
1
40
Q
20 Q
0
Q 20
h
10 Q hm

0 0
BC 30 15 BC 30 15 21 22 23 24 25

Figure 11.2: Heat demand and recovered energy (left) and efficiencies (right) for the base case (BC)
and the buildings with a nominal energy demand for heating of 30 kW h/m2 a (30) and
15 kW h/m2 a (15)

An increase of 4 with 8.1 and 17.6 percentage points is observed, caused by an increase of
Q0 Q and a decrease of Q0 Qnv , with increasing insulation. This means the increase of the
heat demand caused by ventilation the house, Q0 Qnv , is reduced more by a heat exchanger,
when the building is better insulated. Without a heat exchanger, this increase is almost the
same for the three buildings: 31.58 kW h/m2 , 31.81 kW h/m2 and 30.34 kW h/m2 .

11.3 Ventilation flow rate


It is clear that a decrease of the ventilation flow rate, with less extraction of warmer indoor
air and less supply of colder outdoor air, will lower the heat demand of the building. The heat
demands and the recovered energy for the buildings with lower ventilation air flow rates are
shown on the left graph of Figure 11.3 and the influence on the efficiencies on the right graph
of the same figure. When the terms medium and low are used, this means the ventilation flow
rates have been reduced with respectively 2/3 and 1/3, compared to the base case.
Except for Qnv , which is independent of the ventilation system, all components on the left
side of Figure 11.3 decrease. A smaller ventilation flow rate means less redistribution of cold
air from non-heated rooms to heated rooms. This is seen by the decrease in Q1 Qnv . The
other components, Q and Q0 decrease because lower flow rates mean a smaller amount of energy
exhausted to the atmosphere. That less energy is extracted and consequently released in the
atmosphere, can also be seen by the decreasing values of Qh and Qhm
Opposed to the discussion of the previous parameters, 1 and 3 are no longer constant,
although the same heat exchanger is used. Because the air flow rate is decreased, the effective-
ness of the heat exchanger is higher, in accordance with the -NTU correlations, explained in
Section 6.1.3. With these equations, the effectiveness, which was 75% at nominal flow rates, can
be calculated with the reduced flow rates. This gives an effectiveness of 81.8% for a reduction
of the flow rates with a factor 2/3 and an effectiveness of 90.0% for a reduction with a factor
1/3. Because of the linear relationship between the effectiveness and 1 and 3 , explained in
the discussion of the base case in Chapter 10, these are also the values of 1 and 3 seen on
Figure 11.3.
A similar value of 2 , the fraction of the recovered heat that is supplied usefully, is observed
when varying the flow rate. A lower flow rate means less extracted heat and a smaller value of
CHAPTER 11. INFLUENCE OF PARAMETERS ON RECOVERY PERFORMANCE 54

50 100
Q nv
Q1
40 80
Q
Energy, kWh/m2

Q0

Efficiency, %
30 Qh 60
Q
hm

20 40

10 20 Base case
Medium setting
Low setting
0 0
BC M L BC M L 21 22 23 24 25

Figure 11.3: Heat demand and recovered energy (left) and efficiencies (right) for the base case (BC)
and the buildings with the ventilation system used on medium (M) and low (L)

Qh . But less extracted heat also means a higher indoor air temperature and so less recovered
energy that is supplied usefully (Q0 Q decreases). This balances out, keeping 2 approximately
constant.
The increasing value of 4 indicates that with decreasing flow rates, a higher fraction of the
extra heat required by ventilating the building, can be recovered. This is caused by a decreasing
value of Q0 Q but an proportionally more increased value of Q0 Qnv , because there is less
redistribution of the air over the rooms.
Finally 5 increases with 4.38 and 10.07 percentage points for the lower flow rates. This is caused
by the similar values of 2 and the higher values of 3 .

With the reduced flow rates, a reduction in heat demand, Q, from 30.67 kW h/m2 for the
base case to 23.38 kW h/m2 and 17.96 kW h/m2 on medium and low setting is observed. This
can be attributed to the lower flow rates itself as well to the increased effectiveness of the heat
exchanger (caused by the lower flow rates). That Q is lower is caused by a smaller value of
Q1 Qnv and a smaller value of Q Q1 , since Q = Qnv + (Q1 Qnv ) + (Q Q1 ) and Qnv
doesnt change with a changing ventilation flow rate. The decrease in Q1 Qnv is fully the
result of the lower flow rates itself and is not influenced by the -correlations, since the effec-
tiveness is not important for both heat demands. The second part, Q Q1 , can be written as
(1 1 ) (Q0 Q1 ) (1 S ) (Q0 Q1 ), which shows that this is again caused by both the
increased effectiveness (1 S ) and the lower flow rates itself (Q0 Q1 is again independent of
the effectiveness). Using this equation and the current values of Q0 and Q1 , the heat demand
that the building would have with a fixed effectiveness of 75%, regardless of the flow rates, can
be calculated. These values are 24.63 kW h/m2 with the ventilation system on medium setting
and 19.20 kW h/m2 on low setting. Simulations are done to verify these calculations where the
effectiveness was fixed at 75%, regardless of the flow rates, and the found heat demands are
24.48 kW h/m2 and 19.12 kW h/m2 , a difference of maximum 0.61% with the calculated values.

Important to note is that although the recovery efficiencies are highest for the lowest flow
rates, the amount of energy that is usefully recovered, Q0 Q, is highest for the base case.
Q0 Q has a value of 17.74 kW h/m2 for the base case, 13.22 kW h/m2 for a reduction of the
flow rates with a factor 2/3 and 7.14 kW h/m2 for a reduction of the flow rates with a factor
1/3. The heat demand for the building with the lowest flow rates without heat exchanger is still
CHAPTER 11. INFLUENCE OF PARAMETERS ON RECOVERY PERFORMANCE 55

even lower than the heat demand of the building with most ventilation with a heat exchanger.

Please also note that when comparing the cases with different flow rates, that by decreasing
the ventilation air flow rates, the indoor air quality deteriorates. The average CO2 -concentration
in the occupied rooms increases from 667 ppm for the base case, to 784 ppm for the medium
setting and 1025 ppm for the low setting.

11.4 Demand-controlled ventilation


Since with the demand-controlled ventilation strategy used in this study (explained in Sec-
tion 6.1.4), the ventilation air flow rates are alternated between 100% and 10% of the nominal
flow, this case has the same trends as the case discussion the lower flow rates (Section 11.3).
The results are given in Figure 11.4.
All heat demands decrease, except for Qnv which is independent of the ventilation flow rates.
Also the energy recovered by the heat exchanger decreases because of the lower flow rates.
Since the lower flow rates increase the effectiveness of the heat exchanger (-NTU correla-
tions) increases with causes higher values of 1 and 3 . 2 stay approximately the same, which
gives, multiplied with 3 , an higher value of 5 . Also the same as with the previous discussion
is the higher value of 4 , caused by a decrease in Q0 Q and a more important decrease in
Q0 Qnv , because less energy is redistributed over the rooms.
This lower heat demand and higher efficiency come at the price of a worse indoor air quality.
With the CO2 -based DCV strategy, the average CO2 -level in the occupied rooms is 827 ppm,
which is 150 ppm more than for the base case.

50 100
Q nv Base case
Q1 DCV
40 80
Q
Q0
Energy, kWh/m2

Efficiency, %

30 Qh 60
Q hm

20 40

10 20

0 0
BC DCV BC DCV 21 22 23 24 25

Figure 11.4: Heat demand and recovered energy (left) and efficiencies (right) for the base case (BC)
and the case which demand-controlled ventilation (DCV)

11.5 Heat exchanger effectiveness


It was already mentioned that 1 and 3 vary linearly with the sensible effectiveness of the heat
exchanger (Figure 10.5). The influence of the sensible effectiveness on the heat demand and
the efficiency can be seen in Figure 11.5. To become these results, sensible effectiveness was
increased to 80% and 85%. On the left, it can be seen that only heat demand Q changes. Q0 ,
Q1 and Qnv are independent of the sensible effectiveness of the heat exchanger. The recovered
energy, Qh , increases with increasing effectiveness, as expected. Also Qhm slightly increases.
This last increase is caused by the higher indoor temperatures, since the supplied air is warmer.
CHAPTER 11. INFLUENCE OF PARAMETERS ON RECOVERY PERFORMANCE 56

50 100
Base case
0 =80%
S
40 80 0 S =85%
Energy, kWh/m2

Efficiency, %
30 60
Q nv
20 Q1 40
Q
Q0
10 20
Qh
Q hm
0 0
BC 80% 85% BC 80% 85% 21 22 23 24 25

Figure 11.5: Efficiencies for the base case and with increased sensible effectiveness of the heat exchanger

1 and 3 increase as expected to 80% and 85%. 2 , the fraction of the heat recovered by
the heat exchanger that is usefully supplied to the rooms is almost constant at slightly less than
50%. These trends from 2 and 3 explain the increase of 5 .
Finally, an increase of 4 with 3.74 and 7.45 percentage points compared to the base case is
observed. This efficiency is higher, since Q0 Q increases (decreasing value of Q with higher
effectiveness of the heat exchanger), while Q0 Qnv is the same for the base case as with a
sensible effectiveness of 80% or 85% (Q0 and Qnv are both independent of the effectiveness of
the heat exchanger).

11.6 Heating and occupancy pattern


For this study, a heating strategy is chosen where a room is only heated when occupied, as
mentioned in Section 6.2. The occupancy is divided over five (heated) rooms and the occupancy
profiles are explained in Section 7.2. In the base case, the building has a low occupancy. With
the second occupancy profile, the members of the household are more at home than with the
base case. This means more hours the rooms need to be heated, since they are only heated when
somebody is present. On the other hand, since people produce heat, a higher occupancy also
means more internal heat gains. As can be seen on Figure 11.6 (left), this first factor prevails,
because all heat demands increase. Also the recovered energy and the maximally recoverable
energy are higher when the people are mostly at home, since the room temperatures during the
heating season are higher on average.
Since 1 and 3 are the same as the effectiveness of the heat exchanger, these efficiencies are
approximately 75% for both occupancy profiles. With a higher presence of people, the recovered
energy is less supplied when no heating is necessary, increasing the useful fraction. Therefore 2
and 5 are higher when the building is more often occupied.

Because 2 and 5 improved with increasing occupancy, simulations were done with two
extreme occupancy profiles to see if this trend continues. First the five heated rooms, over
which the occupancy was originally divided, are continuously heated regardless of the occupancy.
Secondly all rooms, also the originally non-heated rooms, are continuously heated. As expected
the heat demand increases tremendously, from 30.67 kW h/m2 for the base to 49.21 kW h/m2
and to 57.37 kW h/m2 for the two extreme occupancy profiles, but also the values of 2 and 5
increase, as can be seen in Figure 11.7. Since so much heating is required, less energy is not
supplied usefully.
CHAPTER 11. INFLUENCE OF PARAMETERS ON RECOVERY PERFORMANCE 57

60 100
Base case
50 Mostly at home
80
Energy, kWh/m2

Efficiency, %
40
60
30 Q nv
Q1 40
20 Q
Q0
20
10 Qh
Q hm
0 0
BC MAH BC MAH 21 22 23 24 25

Figure 11.6: Heat demand and recovered energy (left) and efficiencies (right) for the base case (BC)
and the mostly at home occupancy pattern (MAH)

100
Base case
Always heated
80 Every room heated
Efficiency, %

60

40

20

0
22 25

Figure 11.7: 2 and 5 compared to the base case with two extreme heating strategies
CHAPTER 11. INFLUENCE OF PARAMETERS ON RECOVERY PERFORMANCE 58

11.7 Comfort level


The comfort level is adapted by changing the temperatures to which the rooms are heated
(Section 6.2). For the base case, this is the neutral temperature and for the two variants studied
here, this are the lower and upper limit of the 10 P P D range around this neutral temperature,
for low and high comfort level respectively. In Figure 11.8, the results are shown.
On the left graph, it can be seen that, as expected, every heat demand and the recovered
energy decreases with the lower set-point temperatures and increases when a higher comfort
level is demanded.
When the comfort level is low and the set-point temperatures decrease, the heated rooms
will sooner reach their demanded temperature and recovered energy that is then supplied no
longer directly contributes to reducing the heat demand. Therefore less of the recovered energy
is supplied usefully to the building with this low comfort level (and vice versa for the high
comfort level), as can be seen from 2 . This trend is also seen with 5 , since 1 and 3 are again
constant and 5 = 2 3 . 4 is almost independent of the comfort level, since the values are
approximately the same.

80 100
Base case
70 Low comfort
High comfort
80
60
Energy, kWh/m2

Efficiency, %

50
60

40
Q nv
40
30 Q1
Q
20 Q
0 20
Q
10 h
Q hm

0 0
BC LC HC BC LC HC 21 22 23 24 25

Figure 11.8: Heat demand and recovered energy (left) and efficiencies (right) for the base case (BC),
the low comfort case (LC) and the high comfort case (HC)

11.8 Summary
The results of the current chapter are summarised in Figure 11.9. On this figure, the bars (left
y-axis) give the fraction of the maximally recoverable heat that was supplied usefully to the
building, 5 . On the right y-axis, the amount of energy that was saved by installing a heat
exchanger, Q0 Q, is shown. The two horizontal lines indicate the efficiency and heat demand
reduction of the base case.

The flow rate has the biggest influence on the efficiency. A maximum increase with 10.1 per-
centage points compared to the base case was achieved when the flow rate was reduced with a
factor 1/3. Also the occupancy pattern has a high influence on the efficiency, since the efficiency
for case with a high occupancy is 8.93 percentage points higher than the base case with its low
occupancy profile. The type of building and its nominal energy demand have a smaller effect
CHAPTER 11. INFLUENCE OF PARAMETERS ON RECOVERY PERFORMANCE 59

on the efficiency than e.g. flow rate. Buying a heat exchanger with a higher effectiveness does
increase the efficiency of the energy recovery, but an increase of the effectiveness with 10 per-
centage points does, for this case, only mean an increase in efficiency with 4.21 percentage points.

Next to the efficiency it is also interesting to look at the reduction of the heat demand by
installing a heat exchanger between the two air streams. It was mentioned that the efficiency was
the highest for the lowest flow rates, but the energy that is recovered on these lower settings is
much lower than on the nominal setting (also for the DCV case, since this is actually a reduction
of the average flow rates). This could have an impact on e.g. payback time. An increase of
Q0 Q is observed for the better insulated houses, the cases with a better heat exchanger, the
higher occupancy and the higher comfort. The type of building does not seem to have a big
influence on the amount of energy that is supplied usefully to the building.

100 30

Heat demand reduction, kWh/m 2


25
Q 0 -Q 25
80
Efficiency, %

20
60
15
40
10

20
5

0 0
cy
V
se

t
d

ate

ate

t
%

%
d

for
for
h/m 2

h/m 2
he

ce

DC

80

85

n
ca

wr

wr

om
om
pa
rra
tac

0=

0=
se

ccu
flo

flo

hc
wc
Te

kW

kW
-de
Ba

2/3

1/3

Hig
ho

Lo
mi

30

15
Se

Hig

Figure 11.9: Heat recovery efficiency and heat demand reduction by the heat exchanger for the base
case and the variations
Chapter 12

Comparison with type C ventilation


systems

Up till now, all buildings studied had both mechanical supply and mechanical extraction (type
D ventilation system). This chapter compares the performance of these buildings with houses
with natural supply and mechanical extraction (type C ventilation system), which do not have
a heat exchanger.
In the first part of this chapter, the buildings with a ventilation system that is always on are
studied. The second part discusses buildings with demand-controlled ventilation.

12.1 No demand-controlled ventilation


The required supply and extraction flow rates for the type D ventilation systems and the required
extraction flow rates for the type C ventilation system are given in Table 6.1. The buildings
with a type D ventilation system require higher exhaust flow rates. This is because the supply
and exhaust flow rates need to be balanced. The natural supply to the so-called dry rooms in
the buildings with a type C ventilation system is done with flow controllers. These are described
in Section 6.1.2.

The heat demand of the buildings with these flow rates (continuously, no DCV) are calculated
with the method explained in Section 9.1. For the building with the type D ventilation system
(base case), the heat demand is 30.67 kW h/m2 . The same house as the base case, but with
natural supply has the advantage of the lower required ventilation flow rates, but the naturally
supplied air is not preheated by a heat exchanger. The resulting heat demand is 24.12 kW h/m2
or a decrease of 6.55 kW h/m2 . This energy saving should be weighed against the indoor air
quality. The average concentration of CO2 in the occupied rooms of the house with a type C
ventilation system is 538 ppm higher compared to its type D counterpart (1215 ppm compared
to 667 ppm).

To examine the differences between these two types of ventilation system more deeply, the
ventilation flow rates are varied for each case and the energy demand is visualised against the
IAQ. The results for the detached building are given in Figure 12.1. Five flow rates are compared,
of which the lowest flow rates (175 m3 /h extracted in total) correspond to the flow rates required
with the type C ventilation system. The flow rates are then linearly varied in three steps to
the flow rates required with the type D ventilation system (361 m3 /h extracted in total) and
extrapolated one step further.
The discussion of this figure is first done for the three versions of the building with the HRV
system (marked with a circle). They differ in level of insulation and airtightness and are labelled

60
CHAPTER 12. COMPARISON WITH TYPE C VENTILATION SYSTEMS 61

by their nominal energy demand for heating: 15 kW h/m2 a, 30 kW h/m2 a and 60 kW h/m2 a.
Then the same buildings, but with the type C ventilation system (marked with a triangle) are
discussed mutually. Finally, both types of ventilation systems are compared with each other.

40 423
361
Heat demand, kWh/m2
299 C, 15 kWh/m 2a
30 237
C, 30 kWh/m 2a
175
C, 60 kWh/m 2a
2
20 D, 15 kWh/m a
D, 30 kWh/m 2a
D, 60 kWh/m 2a
10

0
600 800 1000 1200
CO2-concentration, ppm

Figure 12.1: Heat demand versus indoor air quality with varying ventilation flow rates (in m3 /h) for
the different versions of the detached building

For the buildings with the mechanical supply of fresh air, the heat demand increases with
increasing flow rates and the IAQ improves. This was expected, because higher flow rates mean
a more frequent replacement of polluted indoor air with fresh outdoor air, so a lower CO2 -
concentration, but more heat required to increase the temperature of the supplied air to room
temperature. The heat demand decreases with increasing level of insulation and airtightness,
because there are less transmission and infiltration losses. The CO2 -demand does not vary for
the three versions of the building with the same flow rates, because in each corresponding room,
the same amount of air is extracted or supplied.
The same trend of increasing heat demand and improving IAQ with increasing flow rates is
observed for the buildings with the type C ventilation system. Also the decreasing heat demand
with increasing insulation and airtightness can be seen when comparing the three version of the
building mutually. What has changed, compared to the buildings with the type D ventilation
system, is that for the same flow rates, the CO2 -concentration drops when the airtightness is
improved. This seems strange, because with a leaky building, one would expect a better indoor
air quality. To investigate what causes this observation, the average CO2 -concentration in the
rooms of these buildings with the lowest flow rates is given by Figure 12.2. All corresponding
rooms seem to have approximately the same IAQ for the house, independent of the nominal
energy demand for heating, except for bedroom 1 (the only occupied bedroom).
For a possible explanation, it is noticed from the floor plan of the first floor of the detached
building (see Figure 5.1 or Appendix A) that the only room with extraction of air on this
floor is the bathroom. When a fan extracts air from this room, the pressure in this room and
consequently the pressure on the entire first floor drops. It is the pressure drop in the bathroom
that enables fresh air to be supplied to the bathroom, either from the hall or through cracks, and
the pressure drop of the first floor that supplies the bedrooms with fresh air, either through the
flow controllers or through cracks. For the building with the worst airtightness, most air enters
the bathroom through cracks in the external wall, which means less air needs to be supplied
from the hall, the underpressure in the bedrooms is smaller and less air is supplied through the
flow controller of these bedrooms. From the simulations, it was indeed seen, that the average
air flow rate supplied through the flow controller in bedroom 1 of the building with an energy
CHAPTER 12. COMPARISON WITH TYPE C VENTILATION SYSTEMS 62

1500
15 kWh/m 2a

CO2-concentration, ppm
30 kWh/m 2a
1000 60 kWh/m 2a

500

3
ll
dy
e

ilet
en

m
m

Ha
rag

oo
roo

Stu
ch

To

roo

roo

roo

thr
Sto
Kit
ing

Ba
Be

Be

Be
Liv

Figure 12.2: Average CO2 -concentration in the rooms of the three versions of the detached building
with the type C ventilation system and the lowest flow rates

demand for heating of 15 kW h/m2 a was 1.89 times higher than for the building with an energy
demand for heating of 60 kW h/m2 a. The phenomenon of the worse IAQ is not observed for the
other dry rooms (living room, study and the two other bedrooms). For the two other bedrooms,
it is simply because they are not occupied and there is no generation of CO2 . The living room
and the study are sometimes occupied but for far less long periods than the bedroom which is
occupied for several hours during the night. They are also located on the ground floor, where
there is extraction in the toilet, kitchen and storage room and so a higher extraction flow rate
in total.
The results of the buildings with type D and type C ventilation systems have now been
discussed separately. When comparing the two situations, it can be seen that the heat demand
in the building with mechanical supply is always lower for the same ventilation air flow rate
than for the corresponding building with natural supply.
The same is valid for the IAQ, although the improvement gets smaller with increasing flow rates
and with increasing airtightness because of the phenomenon described with the discussion of
the buildings with the type C ventilation system.

408
40 350 40 418
357
Heat demand, kWh/m2

Heat demand, kWh/m2

292
296
233
30 30 236
175
175
2
C, 15 kWh/m a
20 2
20
C, 30 kWh/m a
2
C, 60 kWh/m a
10 D, 15 kWh/m 2a 10
D, 30 kWh/m 2a
D, 60 kWh/m 2a
0 0
600 800 1000 1200 600 800 1000 1200
CO -concentration, ppm CO -concentration, ppm
2 2

Figure 12.3: Heat demand versus indoor air quality with varying ventilation flow rates (in m3 /h) for
the different versions of the semi-detached and the terraced building

In Figure 12.3, the results for the semi-detached and the terraced buildings are given. The
same trends are visible as with the discussion of the terraced building. Since the semi-detached
CHAPTER 12. COMPARISON WITH TYPE C VENTILATION SYSTEMS 63

building has one wall shared with an adjacent neighbour and the terraced building has two,
they have a smaller external wall area. Because of this, the total air flow rate through cracks
becomes smaller relative to the air flow through the flow controllers, for the buildings with the
type C ventilation system. This is the reason the phenomenon of improving IAQ with increas-
ing airtightness becomes smaller compared to the detached buildings, which has no adjacent
neighbours.

12.2 Demand-controlled ventilation


When the ventilation is only used when its necessary, the term demand-controlled ventilation is
used. The effects of DCV on the energy recovery performance with a type D ventilation system
were already discussed in Section 11.4. The heat demand decreases and the efficiency increases,
but the IAQ deteriorates. Here, the type C ventilation system with DCV, commercially often
called C+, is discussed. The same CO2 -based DCV strategy is used as explained in Section 6.1.4
and used in Section 11.4. The ventilation is turned on when the maximum CO2 -concentration
in the building exceeds 1000 ppm and is reduced to 10% of the nominal flow rates when this
concentration drops below 900 ppm.

The comparison of the C and C+ type ventilation system is done based on Figure 12.4. Since
the DCV strategy basically means lower average ventilation flow rates, the observed trends are a
decrease in heat demand, combined with a worse IAQ. The differences in heat demand and IAQ
for the corresponding cases with and without DCV are the smallest for the buildings with the
smallest flow rates. The reason is that for these buildings, the difference between the average flow
rates with and without DCV are much smaller than for the buildings with the highest flow rates.

Now the comparison is made between the type C+, without heat recovery and type D, with
heat recovery. As can be seen in Figure 12.5, the IAQ is always better for the cases with the
type D ventilation system than the type C+ ventilation system. The same is not valid for the
heat demand. For the building with highest energy demand for heating (60 kW h/m2 a), the
heat demand for the three highest flow rates is lowest for the type C+ system, but for the two
lowest flow rates, it is lowest for the type D system. Apparently, the saving by using the DCV
strategy are most significant for the higher flow rates, because there, the difference between the
on- and off-state is the greatest. As was discussed in Section 11.2, the efficiency of the energy
recovery increases when the building is better insulated and more airtight. This causes the
decrease in difference in heat demand between the type C+ and type D for the lower nominal
energy demands for heating. For the best insulated house (15 kW h/m2 a), the heat demand is
only smaller for the type C+ ventilation system at the highest flow rate. To illustrate the better
performance of the type D ventilation system over the type C+ ventilation system, one can see
that the buildings with the type D ventilation system on the second lowest flow rates (237 m3 /h
extracted in total) not only have a better IAQ than the corresponding buildings with the type
C+ ventilation system even at the highest flow rates, but also a lower heat demand than the
type C+ ventilation system at the lowest flow rates.
CHAPTER 12. COMPARISON WITH TYPE C VENTILATION SYSTEMS 64

40 423
361
Heat demand, kWh/m2

299
30 237 C, 15 kWh/m 2a

175 C, 30 kWh/m 2a
C, 60 kWh/m 2a
20 C+, 15 kWh/m2a
C+, 30 kWh/m2a
C+, 60 kWh/m2a
10

0
600 800 1000 1200
CO -concentration, ppm
2

Figure 12.4: Heat demand versus indoor air quality for the type C and C+ ventilation systems in the
detached building at different flow rates

40
2
Heat demand, kWh/m

30 423 C+, 15 kWh/m2a


361
299 C+, 30 kWh/m2a
237
175
C+, 60 kWh/m2a
20 D, 15 kWh/m 2a
D, 30 kWh/m 2a
D, 60 kWh/m 2a
10

0
600 800 1000 1200
CO2-concentration, ppm

Figure 12.5: Heat demand versus indoor air quality for the type C+ and D ventilation systems in the
detached building at different flow rates
Chapter 13

Conclusion

In this study, the performance of heat recovery systems in buildings with a low energy demand
for heating is investigated. The performance is indicated by the efficiency and the size of the
reduction in heat demand, compared to the same building without heat exchanger. The influ-
ence of the type of building, the level of insulation and airtightness, the ventilation flow rates,
the ventilation strategy, the heat exchanger effectiveness, the occupancy and the comfort de-
mand on the performance are examined. Furthermore, a comparison is made with buildings with
mechanical extraction of polluted air but natural supply of fresh air (type C ventilation system).

To compare the different cases, the buildings are simulated with the software package TRN-
SYS. The internal heat gains and the occupancy profiles are made to represent a household with
two members. The weather data is based on the climate in Uccle, Belgium.

By installing a heat exchanger, the heat demand decreases with 17.74 kW h/m2 or 36.65%
for the base case. The efficiency of the energy recovery, i.e. the fraction of the heat extracted
by the ventilation system that is recovered and supplied to the building usefully, is 37.56%.

Beside the fact that a terraced and semi-detached house need less insulation than a detached
house to achieve the same energy demand, no real conclusion regarding the recovery performance
could be made for the type of building.
For a more insulated and airtight building, more energy is saved, but the efficiency is approxi-
mately the same.
Lowering the ventilation flow rates increases the efficiency, but lowers the reduction in heat
demand and impairs the IAQ. The same is valid for a demand-controlled ventilation system,
since this is basically a reduction of the average flow rate.
The increase of the effectiveness of the heat exchanger from 75% to 85% means here a decrease
of the heat demand with 7.66% and an increase of the recovery efficiency with 4.21 percentage
points.
More occupancy and a higher demanded comfort level both mean a higher heat demand, but also
a higher efficiency of the recovery and more energy saved compared to the same cases without
heat exchanger.

When comparing the type D ventilation system with the type C ventilation system, it is
seen that a building with mechanical supply will always have a better IAQ than a building
with natural supply. Also the heat demand is lower for the buildings equipped with a type D
ventilation system, especially for buildings with more insulation and a better airtightness.
For high ventilation flow rates, a building with a demand-controlled type C ventilation system,
referred to as C+, has a lower heat demand than this building with a type D ventilation system
and the same flow rates, but a higher average CO2 -concentration in the occupied rooms. And

65
CHAPTER 13. CONCLUSION 66

when the flow rates of this building with the type D ventilation system are reduced, the heat
demand drops below that of a type C+ ventilation system, still achieving a better indoor air
quality.
Bibliography

[1] European Commission. Europe 2020: A Strategy for Smart, Sustainable and Inclusive
Growth: Communication from the Commission. Publications Office of the European Union,
2010.

[2] Directive EU. 2010/30/EU of the European Parliament and of the Council of 21 May 2008
on the indication by labelling and standard product information of the consumption of
energy and other resources by energy-related products. Official Journal of the European
Union, 2010.

[3] Egidijus Juodis. Extracted ventilation air heat recovery efficiency as a function of a build-
ings thermal properties. Energy and Buildings, 38(6):568573, 2006.

[4] Augustino Binamu and Ralf Lindberg. The impact of air tightness of the building envelope
on the efficiency of ventilation systems with heat recovery. 2001.

[5] C.-A Roulet, F.D Heidt, F Foradini, and M.-C Pibiri. Real heat recovery with air handling
units. Energy and Buildings, 33(5):495502, 2001.

[6] Younness El Fouih, Pascal Stabat, Philippe Riviere, Phuong Hoang, and Valerie Archam-
bault. Adequacy of air-to-air heat recovery ventilation system applied in low energy build-
ings. Energy and Buildings, 54:2939, 2012.

[7] Andy P Jones. Indoor air quality and health. Atmospheric environment, 33(28):45354564,
1999.

[8] J. Hansen, M. Sato, P. Kharecha, D. Beerling, R. Berner, V. Masson-Delmotte, M. Pa-


gani, M. Raymo, D. L. Royer, and J. C. Zachos. Target atmospheric CO2: Where should
humanity aim? Open Atmos. Science Journal, 2:217231, 2008.

[9] OA Seppanen, WJ Fisk, and MJ Mendell. Association of ventilation rates and co2 concen-
trations with health andother responses in commercial and institutional buildings. Indoor
air, 9(4):226252, 1999.

[10] World Health Organization et al. Air quality guidelines for europe. 2000.

[11] Antigifcentrum. CO2 -INTOXICATIES 2014. http://www.antigifcentrum.be/


koolstofmonoxide/co-vergiftiging. Accessed: 07-11-2015.

[12] Marco Maroni, Bernd Seifert, and Thomas Lindvall. Indoor air quality: a comprehensive
reference book. Elsevier, 1995.

[13] L Mlhave, Bodi Bach, and Ole F Pedersen. Human reactions to low concentrations of
volatile organic compounds. Environment International, 12(1):167175, 1986.

[14] Vlaanderen. Ventilatie - energiesparen. http://www.energiesparen.be/ventilatie, .


Accessed: 07-11-2015.

67
BIBLIOGRAPHY 68

[15] BIN. NBN D 50-001: Ventilatievoorzieningen in woongebouwen, 1991.

[16] Richard Aynsley. Estimating summer wind driven natural ventilation potential for indoor
thermal comfort. Journal of Wind Engineering and Industrial Aerodynamics, 83(1):515
525, 1999.

[17] Gian Vincenzo Fracastoro and Matteo Serraino. Energy analyses of buildings equipped
with exhaust air heat pumps (eahp). Energy and Buildings, 42(8):12831289, 2010.

[18] Mike Schell and Dan Inthout. Demand Control Ventilation Using CO2 . ASHRAE journal,
2001.

[19] Nathan Van Den Bossche, Arnold Janssens, N Heijmans, and P Wouters. Performance
evaluation of humidity controlled ventilation strategies in residential buildings. Thermal
performance of the exterior envelopes of whole buildings X. Clearwater, page 7, 2007.

[20] Toke Rammer Nielsen and Christian Drivsholm. Energy efficient demand controlled venti-
lation in single family houses. Energy and Buildings, 42(11):19951998, 2010.

[21] Tao Lu, Xiaoshu Lu, and Martti Viljanen. A novel and dynamic demand-controlled venti-
lation strategy for CO2 control and energy saving in buildings. Energy and Buildings, 43
(9):24992508, 2011.

[22] Michael J. Brandemuehl and James E. Braun. Impact of demand-controlled and economizer
ventilation strategies on energy use in buildings. ASHRAE Transactions, 105, 1999.

[23] C.Y.H. Chao and J.S. Hu. Development of a dual-mode demand control ventilation strategy
for indoor air quality control and energy saving. Building and Environment, 39(4):385397,
2004.

[24] Stanley A Mumma. Demand Controlled Ventilation using DOAS. ASHRAE IAQ Applica-
tions, pages 1921, 2002.

[25] Mike B Schell, Stephen C Turner, and R Omar Shim. Application of CO2 -Based Demand-
Controlled Ventilation Using ASHRAE Standard 62: Optimizing Energy Use and Venti-
lation. Transactions - American Society of Heating, Refrigerating and Air-Conditioning
Engineers, 104:12131225, 1998.

[26] A Mardiana-Idayu and SB Riffat. Review on heat recovery technologies for building appli-
cations. Renewable and Sustainable Energy Reviews, 16(2):12411255, 2012.

[27] Paul Heat Exchanger - Standard heat exchangers HRV - Product infor-
mation. URL http://paul-waermetauscher.de/en/product-information/
standard-heat-exchangers-hrv.html. accessed: 19-11-2015.

[28] Ulrich Sander and Harald Janssen. Industrial application of vapour permeation. Journal
of Membrane Science, 61:113129, 1991.

[29] M. Nasif, R. AL-Waked, G. Morrison, and M. Behnia. Membrane heat exchanger in HVAC
energy recovery systems, systems energy analysis. Energy and Buildings, 42(10):18331840,
2010.

[30] Jingchun Min and Ming Su. Performance analysis of a membrane-based energy recovery
ventilator: Effects of outdoor air state. Applied Thermal Engineering, 31(17-18):40364043,
2011.
BIBLIOGRAPHY 69

[31] J. L. Niu and L. Z. Zhang. Membrane-based Enthalpy Exchanger: Material considerations


and clarification of moisture resistance. Journal of Membrane Science, 189(2):179191,
2001.

[32] Sadik Kakac, Hongtan Liu, and Anchasa Pramuanjaroenkij. Heat Exchangers: Selection,
Rating, and Thermal Design. CRC press, 2012.

[33] L. Z. Zhang and J. L. Niu. Performance comparisons of desiccant wheels for air dehumidi-
fication and enthalpy recovery. Applied Thermal Engineering, 22:13471367, 2002.

[34] Eurovent. Eurovent Certification. http://www.eurovent-certification.com/index.


php?lg=en. Accessed: 05-21-2015.

[35] HJ Sauer and RH Howell. Promise and potential of air-to-air energy recovery systems.
International Journal of Refrigeration, 4(4):182194, 1981.

[36] William J Fisk and Isaac Turiel. Residential air-to-air heat exchangers: Performance energy
savings, and economics. Energy and Buildings, 5(3):197211, 1983.

[37] J Karlsson, A Roos, and B Karlsson. Building and climate influence on the balance tem-
perature of buildings. Building and environment, 38(1):7581, 2003.

[38] Jelle Laverge, Marc Delghust, Stijn Van de Velde, Tomas De Brauwere, and Arnold
Janssens. Airtightness assessment of newly built single family houses in belgium. In 5th
International BUILDAIR-symposium: Building and ductwork air-tightness. Energie und
Umweltzentrum, 2010.

[39] Mohammad Rasouli, Carey J Simonson, and Robert W Besant. Applicability and optimum
control strategy of energy recovery ventilators in different climatic conditions. Energy and
Buildings, 42(9):13761385, 2010.

[40] Miklos Kassai, Mohammad Rafati Nasr, and Carey J. Simonson. A developed procedure to
predict annual heating energy by heat-and energy recovery technologies in different climate
European countries. Energy and Buildings, 2015.

[41] AF Emery and BL Siegel. Experimental measurements of the effects of frost formation on
heat exchanger performance. In Heat and mass transfer in frost and ice, packed beds, and
environmental discharges. 1990.

[42] Mohammad Rafati Nasr, Melanie Fauchoux, Robert W. Besant, and Carey J. Simonson.
A review of frosting in air-to-air energy exchangers. Renewable and Sustainable Energy
Reviews, 30:538554, 2014.

[43] PT Ninomura and R Bhargava. Heat recovery ventilators in multifamily residences in the
arctic. Technical report, American Society of Heating, Refrigerating and Air-Conditioning
Engineers, Inc., Atlanta, GA (United States), 1995.

[44] S Bilodeau, P Brousseau, M Lacroix, and Y Mercadier. Frost formation in rotary heat and
moisture exchangers. International journal of heat and mass transfer, 42(14):26052619,
1999.

[45] Mohammad Rafati Nasr, Miklos Kassai, Gaoming Ge, and Carey J. Simonson. Evalua-
tion of defrosting methods for air-to-air heat/energy exchangers on energy consumption of
ventilation. Applied Energy, 151:3240, 2015.
BIBLIOGRAPHY 70

[46] Marc Delghust, Jelle Laverge, Arnold Janssens, Charline Van Erck, and Charlotte Taelman.
The influence of user behaviour on energy use in old dwellings: case-study analysis of a
social housing neighbourhood. 5th International Building Physics Conference, Proceedings,
36:809816, 2012.

[47] W Fred Van Raaij and Theo MM Verhallen. Patterns of residential energy behavior. Journal
of economic psychology, 4(1):85106, 1983.

[48] J Paauw, B Roossien, MBC Aries, and O Guerra Santin. Energy pattern generator; under-
standing the effect of user behaviour on energy systems. In 1st European conference energy
efficiency and behaviour, pages 910, 2009.

[49] Olivia Guerra Santin. Behavioural patterns and user profiles related to energy consumption
for heating. Energy and Buildings, 43(10):26622672, 2011.

[50] Reinhard Haas, Hans Auer, and Peter Biermayr. The impact of consumer behavior on
residential energy demand for space heating. Energy and Buildings, 27(2):195205, 1998.

[51] Hugo Hens, Wout Parijs, and Mieke Deurinck. Energy consumption for heating and rebound
effects. Energy and buildings, 42(1):105110, 2010.

[52] Reinhard Haas and Peter Biermayr. The rebound effect for space heating empirical evidence
from austria. Energy policy, 28(6):403410, 2000.

[53] Dorien Aerts, Joeri Minnen, Ignace Glorieux, Ine Wouters, and Filip Descamps. A prob-
abilistic activity model to include realistic occupant behaviour in building simulations.
IBPSA-Canada eSim, 2014.

[54] Ian Richardson, Murray Thomson, David Infield, and Conor Clifford. Domestic electricity
use: A high-resolution energy demand model. Energy and Buildings, 42(10):18781887,
2010.

[55] Runming Yao and Koen Steemers. A method of formulating energy load profile for domestic
buildings in the uk. Energy and Buildings, 37(6):663671, 2005.

[56] Marc Delghust, Arnold Janssens, and Yves Deweerdt. Improving the predictive power of
simplified residential space heating demand models : a field data and model driven study.
2015.

[57] Poul O Fanger et al. Thermal comfort. analysis and applications in environmental engi-
neering. Thermal comfort. Analysis and applications in environmental engineering., 1970.

[58] John W Mitchell and James E Braun. Principles of heating, ventilation, and air conditioning
in buildings. Wiley, 2013.

[59] Richard J de Dear and Gail S Brager. Thermal comfort in naturally ventilated buildings:
revisions to ASHRAE Standard 55. Energy and buildings, 34(6):549561, 2002.

[60] Leen Peeters, Richard de Dear, Jan Hensen, and William Dhaeseleer. Thermal comfort in
residential buildings: Comfort values and scales for building energy simulation. Applied
Energy, 86(5):772780, 2009.

[61] ISO 7730. Ergonomics of the thermal environment Analytical determination and interpre-
tation of thermal comfort using calculation of the PMV and PPD indices and local thermal
comfort criteria. International Organization for Standardisation, Geneva, 2005.

[62] Michael A Humphreys and J Fergus Nicol. The validity of ISO-PMV for predicting comfort
votes in every-day thermal environments. Energy and Buildings, 34(6):667684, 2002.
BIBLIOGRAPHY 71

[63] Gail S. Brager and Richard J. de Dear. Thermal adaptation in the built environment: a
literature review. Energy and Buildings, 27(1):8396, 1998.

[64] Y Tochihara, Y Kimura, I Yadoguchi, and M Nomura. Thermal responses to air tem-
perature before, during and after bathing. In Proceedings of international conference of
environmental ergonomics. San Diego, USA, pages 30913, 1998.

[65] Johannes Theodorus Hendrik Lammers. Human factors, energy conservation, and design
practice. PhD thesis, Technische Hogeschool Eindhoven, 1978.

[66] BW Zingano. A discussion on thermal comfort with reference to bath water temperature
to deduce a midpoint of the thermal comfort temperature zone. Renewable energy, 23(1):
4147, 2001.

[67] CIBSE. Guide A: Environmental design. The Chartered Institution of Building Services
Engineers, London, 2006.

[68] SA Klein et al. A Transient System Simulation Program (TRNSYS 17) Manual. Thermal
Energy System Specialists. Madison, USA, 2010.

[69] FOD Economie Afdeling Statistiek. Kadastrale statistiek van het bestand van de gebouwen,
2015.

[70] Vlaanderen. Belastingvermindering voor lage-energiewoningen,


passiefwoningen en nul-energiewoningen. https://www.
vlaanderen.be/nl/bouwen-wonen-en-energie/bouwen-en-verbouwen/
belastingvermindering-voor-lage-energiewoningen-passiefwoningen-en-.
nul-energiewoningen, . Accessed: 30-04-2016.

[71] Vlaams Energieagentschap. Cijferrapport: 10 jaar energieprestatieregelgeving. Procedures,


resultaten en energetische karakteristieken van het Vlaamse gebouwenbestand - periode 2006
- 2015., 2016.

[72] Heijmans N., Van Den Bossche N., and Janssens A. Berekeningsmethode gelijkwaardigheid
voor innovatieve ventilatiesystemen in het kader van de EPB-regelgeving, toegepast op
SYSTEEM C+ RENSON Ventilation NV. WTCB - Universiteit Gent, 2007.

[73] EPB-productgegevens databank - VENTILATOR EN VENTILATIEGROEP - Erkende


productgegevens (status 1), 2016.

[74] ASHRAE Fundamentals Handbook. American society of heating, refrigerating and air-
conditioning engineers. Inc.: Atlanta, GA, USA, 2009.

[75] Ibrahim Atmaca, Omer Kaynakli, and Abdulvahap Yigit. Effects of radiant temperature
on thermal comfort. Building and Environment, 42(9):32103220, 2007.

[76] Tatiana de Meester, Anne-Francoise Marique, Andre De Herde, and Sigrid Reiter. Im-
pacts of occupant behaviours on residential heating consumption for detached houses in a
temperate climate in the northern part of europe. Energy and Buildings, 57:313323, 2013.

[77] Wolfgang Feist, R Pfluger, B Kaufmann, J Schnieders, and O Kah. Passive house planning
package 2007. Darmstadt: Passive House Institute, 2007.
Appendices

72
Appendix A

Floor plans

In Figure A.1 and Figure A.2, the floor plans of the ground floor and the first floor of the
detached building are shown. The floor plans of the semi-detached and the terraced building are
given in Figure A.3 to Figure A.7. All dimensions on these figures are in mm. For the detached
and the semi-detached buildings, the North is located on the bottom of the figure. The floor
plans of the terraced building have been rotated, and the North is located at the left.

Figure A.1: Floor plan of the ground floor of the detached building

73
APPENDIX A. FLOOR PLANS 74

Figure A.2: Floor plan of the first floor of the detached building

Figure A.3: Floor plan of the ground floor of the semi-detached building
APPENDIX A. FLOOR PLANS 75

Figure A.4: Floor plan of the first floor of the semi-detached building

Figure A.5: Floor plan of the ground floor of the terraced building
APPENDIX A. FLOOR PLANS 76

Figure A.6: Floor plan of the first floor of the terraced building

Figure A.7: Floor plan of the second floor of the terraced building
Appendix B

Building composition

The following tables contain the thickness and heat transfer coefficients of the different compo-
nents of the buildings. It can be seen in each table, that for the floor, the roof and the external
walls, the thickness of the insulation was increased to decrease the energy demand for heating
from 60 kW h/m2 a to 15 kW h/m2 a. When comparing the different building types, it can be seen
that the detached building has the most insulation and the terraced building the least. This
is because the terraced building has two adjacent neighbours, while the detached building has
none. More adjacent neighbours mean a smaller external surface for transmission losses.

Table B.1: Thickness in mm (t) and u-value in W/m2 K (u) of the insulation in the different components
of the different versions of the detached building

Energy demand for heating, kW h/m2 a


15 30 60
Floor t 287 268 220
u 0.215 0.257 0.507
Roof t 285 265 135
u 0.137 0.149 0.320
External wall t 368 349 308
u 0.205 0.243 0.404
Internal wall t 160 160 160
u 3.101 3.101 3.101
Ceiling t 380 380 380
u 1.522 1.522 1.522
Attic floor t 170 170 170
u 0.223 0.223 0.223

77
APPENDIX B. BUILDING COMPOSITION 78

Table B.2: Thickness in mm (t) and u-value in W/m2 K (u) of the insulation in the different components
of the different versions of the semi-detached building

Energy demand for heating, kW h/m2 a


15 30 60
Floor t 270 260 245
u 0.490 0.609 0.959
Roof t 235 208 125
u 0.187 0.214 0.383
External wall t 355 333 294
u 0.230 0.288 0.522
Internal wall t 170 170 170
u 3.018 3.018 3.018
Ceiling t 245 245 245
u 2.155 2.155 2.155
Attic floor t 21 21 21
u 2.867 2.867 2.867

Table B.3: Thickness in mm (t) and u-value in W/m2 K (u) of the insulation in the different components
of the different versions of the terraced building

Energy demand for heating, kW h/m2 a


15 30 60
Floor t 270 245 237
u 0.490 0.959 1.384
Roof t 175 145 82
u 0.236 0.294 0.603
External wall t 319 300 277
u 0.343 0.464 0.809
Internal wall t 170 170 170
u 3.018 3.018 3.018
Ceiling t 240 240 240
u 2.189 2.189 2.189
Attic floor t 21 21 21
u 2.845 2.845 2.845
Flat roof t 294 294 294
u 0.252 0.252 0.252
Appendix C

Calculation of the air mass flow


coefficient for crack modelling

Cracks are modelled by TRNSYS with an exponential leakage characteristic: m = Cs (p)n ,


where m is the air mass flow rate through the crack (kg/s), p the pressure difference over the
crack (P a) and n the air mass flow exponent, chosen to be 0.66 [68].
For the calculation of the Cs -value for the external walls and the roof, the total exfiltration
and infiltration through cracks (mtot ) is first written as the sum of all its components (mw,i and
mr,i for one crack in the wall or the roof):
X
mtot = mw,i + mr,i (C.1)
i

These components are, based on the leakage characteristic, written as:


mw,i = Cs,w Aw,i (p)n (C.2)
n
mr,i = Cs,r Ar,i (p) (C.3)
with Cs,w and Cs,r the air mass flow coefficients for the walls and the roof expressed in kg/(s m2 )
and Aw,i and Aw,r the area of the surface in which the crack is located. When it is assumed that
Cs,r = 2/3 Cs,w [72] and when Aw and Ar are defined as the total wall and roof area, equations
(C.1), (C.2) and (C.3) can be combined to:
mtot = Cs,w (Aw + 2/3 Ar ) (p)n (C.4)
With a pressurization test, the airtightness of a building is determined (typically at 50 P a)
and the leakage flow rate V50 is measured. The airtightness is then indicated by n50 = V50 /V ,
expressed in ach (air changes per hour), where V is the net interior volume of the building.
Since m = V , with 1.2 kg/m3 the air density, the leakage mass flow rate at a pressure
difference of 50 P a can be written as:
m50 = 1.2 V50 = 1.2 V n50 = Cs,w (Aw + 2/3 Ar ) (50)0.66 (C.5)
When the dimensions of a building (V , Aw and Ar ) and its airtightness (n50 ) are known, equation
(C.5) can be used to calculate Cs,w :
1.2 V n50
Cs,w = (C.6)
(50)0.66 (A w + 2/3 Ar ) 2 3600

The factor 2 in the denominator originates from the fact that each surface has two cracks, so
only half the air mass flow rate goes through one crack. Finally, the Cs -values for each crack in
the external walls and the roof, to be used in TRNSYS, is found:
Cs = Cs,w Aw,i (C.7)
Cs = Cs,r Ar,i = 2/3 Cs,w Ar,i (C.8)

79
APPENDIX C. CALCULATION OF THE AIR MASS FLOW COEFFICIENT FOR CRACK MODELLING80

For cracks in internal walls and ceilings, a fixed air flow of 2 m3 /m2 h @ 50 P a is used [72].
This results in a Cs -value of 5.0418 105 kg/m2 s @ 1 P a.
Appendix D

TRNSYS model

The layout of the model in the TRNSYS Simulation Studio is given in Figure D.1. The compo-
nents that generate the outputs are hidden, to keep the figure orderly

In the centre of this figure is the component called Building, indicated with 1. This is
a Type56 - Multi-Zone Building component. This links to the building file that contains all
details on the geometry and composition of the building, e.g. the dimensions of each zone or
the composition of the walls.
In the box labelled, 2, all components are present to model the heat recovery. HEX
is Type667 - Air to Air Heat Recovery component. The latent effectiveness is set to zero,
because this model only has sensible heat recovery and no moisture transfer. The sensible
effectiveness is determined by an equation component, eNTU. This uses the method described
in Section 6.1.3 to calculate the effectiveness using the design effectiveness and the air flow
rates. The component labelled HM is the same type as HEX, but has an effectiveness of
100% and is used to determine the maximum recoverable energy. Also in the same box are
two Type33 - Psychrometrics-components. These calculate the humidity ratio when the percent
relative humidity and temperature are given (or vice versa) and are necessary because Type56
and Type667 each use a different unit to express humidity.
The weather data is generated by the box labelled 3. A Type15-6 - Weather Data Processor
component models the weather based on the climate in Uccle, Belgium and gives e.g. the dry
bulb temperature and the humidity of the air. The LocalWinds component is a Type2260 -
Wind Speed Calculator and determines the local wind speed at different heights, based on the
wind speed at the meteostation and the shear exponent at the location of the building. This
exponent is set to 0.3, which is a rough terrain, corresponding to a small village [72]. The
component named Pressure is used to convert the pressure data from the Type15-6 component
to the right units.
For the calculation of the transmission losses through the floor to the soil, a correct soil
temperature is necessary. This is done with 4, SoilTemp, a Type49 - Ground Coupling Model
component. This component requires a floor plan of the building and the heat transfer into the
building (given by BUILDING) as inputs. With this information, a fully 3-D soil temperature
profile is calculated and the average underfloor surface temperature is given for each room [68].
Next, 5, AirTightness, is an equation component that calculates the air mass flow coeffi-
cients for the modelling of the cracks. It requires the airtightness (6, 3 or 0.6 ach), the volume
and the wall and roof surface of the house and calculates Cs,w and Cs,r with the equations
described in Appendix C.
In 6, the regulating of the fans is done. In FanFactor, is determined what flow rate is
required (1 for the nominal flow rates). The CO2 -based DCV is also done in 6. CO2Control
is a Type2 - AquastatC component and is designed as a thermostat in cooling mode, but in-
stead of turning on cooling when the temperature is to high, it turns on ventilation when the

81
APPENDIX D. TRNSYS MODEL 82

CO2 -concentration is to high. The equation component is used to determine the maximum
concentration in the house.
The internal heat gains are contained in a file that lists for each time step the heat gains in
each room. This file is read by IHG, Type9e - Data Reader. These values are multiplied with
a factor, as explained in Section 7.1, with the equation component called GainFactor.
The occupancy is determined in 8. The data is again read from an external file. The data
reader imports three values every time-step. The first and second value are the number of people
awake and asleep in the house. The last value is a code assigned to each room. This code is
used by the equation component to determine the number of people in each room and the heat
gains by the presence of the people in these rooms.
The last pair of components is used to calculate the set-point temperatures and is located in
the box labelled 9. The comfort temperatures for each room and for every day are written in an
external file and is read by TComf. The equation component SetTemp calculates the set-point
temperatures for each room. These are zero when nobody is present in the room (hence the
link with Occupancy). When somebody is present, the desired air temperature is calculated.
As explained in Section 6.2, comfort is expressed by the desired operative temperature, Top =
Tair +Trad
2 , with Tair the air temperature that is calculated by SetTemp and Trad the mean
radiant temperature given by BUILDING.
APPENDIX D. TRNSYS MODEL 83

2
4
5

9
6
7

Figure D.1: Layout of model in TRNSYS Simulation Studio

You might also like