Sei sulla pagina 1di 148

Design Strategies for

Personalized Ventilation

Ph.D. Thesis

by Radim Čermák

International Centre for Indoor Environment and Energy


Department of Mechanical Engineering
Technical University of Denmark
ISBN: 87–7475–318–5

MEK-I-Ph.D. 04-02
July 2004

International Centre for Indoor Environment and Energy


Department of Mechanical Engineering
Technical University of Denmark
Building 402, 2800 Kgs. Lyngby
Denmark
Table of content

Table of content .................................................................................................................................iii


Preface .................................................................................................................................................. v
Summary ............................................................................................................................................vii
Resumé ................................................................................................................................................ ix
Nomenclature..................................................................................................................................... xi
1. Introduction..................................................................................................................................1
1.1 Background ...........................................................................................................................1
1.2 Principles of personalized ventilation ...............................................................................1
1.3 Design, installation, performance ......................................................................................3
1.4 Contaminants indoors .........................................................................................................7
1.5 Room air distribution...........................................................................................................9
1.5.1 Mixing ventilation ........................................................................................................9
1.5.2 Displacement ventilation...........................................................................................10
1.5.3 Underfloor air distribution........................................................................................14
1.5.4 Control strategies........................................................................................................16
1.6 Interaction of airflows........................................................................................................16
2. Objectives ...................................................................................................................................19
3. Method ........................................................................................................................................21
3.1 Experimental design ..........................................................................................................21
3.2 Air movement room...........................................................................................................24
3.3 Ventilation systems ............................................................................................................25
3.3.1 Personalized ventilation ............................................................................................25
3.3.2 Total-volume ventilation ...........................................................................................26
3.4 Heat and contaminant sources simulation .....................................................................28
3.4.1 Heat sources ................................................................................................................28
3.4.2 Contaminant sources .................................................................................................28
3.5 Measuring Instruments .....................................................................................................30
3.5.1 Breathing thermal manikin .......................................................................................30
3.5.2 Artificial lung ..............................................................................................................31
3.5.3 Tracer-gas analyzer ....................................................................................................32
3.5.4 Low velocity anemometers .......................................................................................33
3.6 Procedure.............................................................................................................................33
3.6.1 Concentration measurement.....................................................................................33
3.6.2 Temperature and velocity measurements...............................................................35
3.7 Criteria for evaluation........................................................................................................36
3.7.1 Contaminant concentration.......................................................................................37
3.7.2 Temperature ................................................................................................................38
3.7.3 Velocity ........................................................................................................................39
3.8 Uncertainty of measurement ............................................................................................39
4. Personalized, mixing and displacement ventilation ..........................................................41
4.1 Objectives.............................................................................................................................41
4.2 Experimental conditions....................................................................................................41
4.3 Visualization of personalized airflow..............................................................................41
4.4 Inhaled air concentration...................................................................................................42

iii
4.5 Inhaled air temperature.....................................................................................................45
4.6 Thermal comfort of seated occupants..............................................................................47
4.7 Contaminant distribution..................................................................................................50
4.7.1 Contaminant concentration profiles ........................................................................50
4.7.2 Inhaled air quality of walking occupants................................................................53
4.8 Temperature distribution ..................................................................................................54
4.9 Velocity distribution ..........................................................................................................55
4.10 Discussion............................................................................................................................56
4.11 Conclusions .........................................................................................................................61
5. Personalized and underfloor ventilation..............................................................................63
5.1 Experiment 1 .......................................................................................................................63
5.1.1 Objectives.....................................................................................................................63
5.1.2 Experimental conditions............................................................................................63
5.1.3 Aerodynamic data of floor diffusers........................................................................63
5.1.4 Inhaled air concentration...........................................................................................66
5.1.5 Inhaled air temperature.............................................................................................67
5.1.6 Contaminant distribution..........................................................................................68
5.1.7 Temperature distribution ..........................................................................................71
5.1.8 Velocity distribution ..................................................................................................72
5.2 Experiment 2 .......................................................................................................................74
5.2.1 Objectives.....................................................................................................................74
5.2.2 Experimental conditions............................................................................................74
5.2.3 Inhaled air concentration...........................................................................................75
5.2.4 Inhaled air temperature.............................................................................................76
5.2.5 Contaminant distribution..........................................................................................77
5.2.6 Temperature distribution ..........................................................................................79
5.2.7 Velocity distribution ..................................................................................................80
5.3 Experiment 3 .......................................................................................................................81
5.3.1 Objectives.....................................................................................................................81
5.3.2 Experimental conditions............................................................................................81
5.3.3 Upward airflow direction..........................................................................................81
5.3.4 Horizontal or downward airflow direction............................................................82
5.4 Discussion............................................................................................................................84
5.5 Conclusions .........................................................................................................................92
6. General discussion....................................................................................................................93
7. Recommendations...................................................................................................................103
8. Conclusions ..............................................................................................................................105
References.........................................................................................................................................107

Appendix A Expression of uncertainty ......................................................................................115


Appendix B Inhaled air concentration with PV, mixing and displacement ventilation ..119
Appendix C Inhaled air temperature with PV, mixing and displacement ventilation.....121
Appendix D Whole-body manikin-based equivalent temperature......................................123
Appendix E Inhaled air concentration with personalized and underfloor ventilation ....125
Appendix F Vertical distribution of active contaminants with underfloor ventilation ...127
Appendix G Risk of airborne infection transmission.............................................................131
Appendix H Intake fraction .........................................................................................................135

iv
Preface

This thesis is the result of a study carried out at the International Centre for Indoor
Environment and Energy, Department of Mechanical Engineering, Technical University of
Denmark from August 2001 to July 2004. The principal supervisor of the study has been
Assoc. Professor Arsen K. Melikov, Ph.D.

I would like to express my gratitude to Professor P. Ole Fanger for inviting me to join the
Centre in 1999, first to carry out my M.Sc. project and later for supporting me to continue
research as a Ph.D. student. I appreciate his encouragement.

I am most grateful to Arsen K. Melikov. It was a great experience being his student. I wish to
thank him for support, numerous discussions, constructive criticism, as well as for sharing
with me his rich experience and inexhaustible optimism. I owe him a great deal.

I would like to thank my colleagues with whom I had the pleasure to cooperate: Luboš
Forejt and Oldřich Kovář for teamwork during the first stage of the experiments and for
carrying out measurements and data analyses; Gabriella Stefanova and Tsvetelina Ivanova
for help during the second stage of the experiments; and Quinfan Zeng for developing a
control system for supplying air for personalized ventilation.

I am grateful to Jan Kaczmarczyk for many discussions on personalized ventilation, and for
his advice during the final stage of writing this thesis.

I would like to extend my gratitude to the scientific, technical and administrative staff at the
Centre. I would like to thank Gunnar Langkilde for help on administrative issues and
Andreas Szekacz for help and advice with experimental facilities. My thanks are due to
Judith Ørting for proofreading this thesis, and to Peter Strøm-Tejsen for translating the
summary of this thesis to Danish.

Last, but certainly not least, I would like to thank my parents and Eva for their support over
the years.

The study has been supported by the Danish Technical Research Council (STVF) as a part of
the research programme of the International Centre for Indoor Environment and Energy
established at the Technical University of Denmark for the period 1998-2007. The Technical
University of Denmark has granted my scholarship.

Kgs. Lyngby, 31. July 2004

Radim Čermák

v
Summary

Personalized ventilation (PV) provides clean air at each workplace. Each occupant is given
an opportunity to improve substantially the quality of inhaled air, and to generate and
control his/her preferred thermal environment. Because occupants may use PV at small
airflow rates and close to isothermal temperatures, a supplementary total-volume
ventilation and air-conditioning system has to be applied in rooms with high heat and/or
pollution load.

In the present study, two types of air terminal device for PV were combined with the most
common total-volume ventilation systems – mixing ventilation, displacement ventilation
and underfloor ventilation. The performance of the combined systems was examined in full-
scale experiments. The criteria for evaluation were air quality and thermal comfort.

A mock-up of a typical office with two identical workplaces was built in a climate chamber.
Two breathing thermal manikins simulating occupants were seated behind each other facing
in the same direction. The air terminal devices tested were a round diffuser mounted on a
movable arm duct, positioned above a computer monitor, and a narrow grill positioned at
the front edge of a desk. Both terminals were adjusted according to the positioning most
often preferred by people. The major difference between the terminals was the direction of
personalized airflow – horizontal/downward in the case of the movable panel and upward
in the case of the desk grill. The arrangement of workplaces did not change during the
whole study, allowing for a direct comparison of the combined systems. Different tracer-
gases were used to simulate the most common contaminant sources indoors. Floor covering,
human bioeffluents and exhaled air were selected. The exposure of occupants was estimated
from the gas concentration measured in the inhalation of the manikins. Temperature of
inhaled air, which is important for the perception of air quality, was measured as well. The
thermal comfort of seated occupants was evaluated based on the heat loss from the
manikins. Furthermore, the distribution of contaminants, temperature and velocity were
measured in order to reveal the airflow pattern in the room. The performance of the PV was
tested at different combinations of personalized airflow rates (0, 7 and 15 L/s). This covered
the range of possible individual patterns of PV use. The total-volume ventilation system was
controlled so that the total cooling capacity of personalized air plus total-volume ventilation
air was constant. The exhaust air temperature was maintained at 26°C.

Results on air quality showed that in rooms with mixing ventilation, PV will always be able
to protect occupants from pollution and thus increase the quality of inhaled air. The mixing
air distribution principle implies that the type of contaminant source (active or passive,
localized or plane) and its location is unimportant. The inhaled air quality of occupants
protected with PV is determined to a large extent by the efficiency of an air terminal device,
and the direction and the rate of personalized airflow. In inhaled air, the temperature was
well correlated with the concentration of the contaminants distributed uniformly in the
room.

In a room with displacement ventilation, PV was shown to improve the inhaled air quality
in regard to a passive and plane contaminant located on the floor. This is the case for carpet,
PVC or linoleum. The use of PV was shown, however, to increase mixing of contaminants
located in the vicinity of the personalized airflow, such as exhaled air and bioeffluents. The

vii
Summary

personalized flow generated large non-uniformities and differences in the distribution of


human-produced contaminants near the workplace. Further from the workplace the non-
uniformities disappeared; however, the concentration of the human-produced contaminants
increased compared to the case without PV. This may decrease the inhaled air quality for
occupants unprotected with PV. The distribution of the human-produced contaminants was
affected by the direction and rate of personalized airflow. Upward personalized flow (desk
grill) was observed to cause a lower transmission of the human-produced contaminants
between workplaces than horizontal/downward flow (movable panel). Personalized airflow
may also affect the distribution of contaminants generated at another workplace. In that
respect, the impact of upward personalized airflow (desk grill) was greater than the impact
of horizontal/downward airflow (movable panel), most probably due to its higher velocity.

The performance of PV in combination with an underfloor air distribution (UFAD) system


with a short throw (up to 0.3 m) was comparable with the performance of PV in combination
with displacement ventilation. UFAD, however, decreased non-uniformities of contaminants
and vertical temperature gradients. Also with UFAD, the use of PV increased mixing of
contaminants located in its vicinity. The experiments confirmed that personalized airflow
directed upward (desk grill) provides a lower transmission of contaminants between
workplaces than personalized airflow directed downward (movable panel). An increase in
the throw diminished gradually the impact of the direction of personalized airflow on the
transmission of contaminants. The impact of the PV direction on the transmission
disappeared when the throw was comparable to the height of the breathing zone.

The analyses indicate that the use of PV in combination with any total-volume ventilation
system could be efficient in protecting occupants even from highly infectious diseases, and
therefore become an alternative or supplement to traditional methods of occupant
protection.

The impact of PV on the thermal environment was very localized, even at high rates of
personalized air (15 L/s). The use of PV at one workplace did not affect the thermal comfort
of the occupant at another workplace. The cooling performance of PV was independent of
the room air distribution generated by a total-volume ventilation system. The cooling
provided by PV was equivalent to decreasing room air temperature by 1-2°C, depending on
the actual combination of PV air terminal, total-volume ventilation principle and workplace.
The ability of PV to affect the air movement in the lower occupied zone was small. In rooms
with thermal stratification (displacement or underfloor ventilation), PV was shown not to
contribute to the removal of heat from the lower zone. This may cause an increase in
temperature in the occupied zone and thus increase the energy consumption of a total-
volume ventilation system. On the other hand, a higher temperature may decrease the risk
of draught and high vertical air temperature difference.

The present study identified that each combination of PV and total-volume ventilation
tested could be applied in practice. Because of an excellent air quality performance and a
low risk of thermal discomfort, a combination of PV and underfloor ventilation with a short
vertical throw, controlled according to a constant air volume strategy, is recommended. If
the control of an airborne transmission of contaminants between occupants has highest
priority, personalized air terminals supplying air upward are preferable. Development of air
terminal devices with a high efficiency and a low ability to promote mixing of contaminants
located in its vicinity is recommended.

viii
Resumé

Personlig ventilation (PV) tilvejebringer ren luft til den enkelte arbejdsplads. Hver bruger
bliver tilbudt en mulighed for at forbedre kvaliteten af indåndingsluften væsentligt, og at
frembringe og kontrollere hans/hendes termiske omgivelser til det foretrukne. Da den en-
kelte bruger muligvis anvender PV med kun en beskeden luftstrøm og tæt ved isotermiske
temperaturer, skal der i rum med en høj varme og/eller forureningsbelastning anvendes et
supplerende ventilations og luftkonditionerings system for rummet.

I nærværende studie blev to forskellige PV indblæsningsterminaler undersøgt i kombination


med de hyppigst anvendte ventilationssystemer hhv. opblandingsventilation, fortræng-
ningsventilation og »underfloor« ventilation. Virkningen af de kombinerede systemer blev
undersøgt ved fuld skala forsøg. Vurderingskriterierne var luftkvalitet og termisk komfort.

En fuld-skala model af et typisk kontor med to identiske arbejdspladser blev opført i et kli-
makammer. To vejrtrækkende termiske mannequiner, simulerende tilstedeværende perso-
ner, blev anbragt siddende bag hinanden med ansigtet vendt i den samme retning. De ind-
blæsningsterminaler, som blev undersøgt, omfattede en rund luftfordeler monteret på en
bevægelig arm placeret over en computer skærm, og en smal rist placeret i arbejdsbordets
forkant. Begge blev justeret i overensstemmelse med den hyppigst foretrukne placering. Den
væsentligste forskel mellem terminalerne var luftstrømningen – horisontalt/nedadrettet i
tilfældet med den bevægelige luftfordeler, og opadrettet i tilfældet med bordristen. Ar-
bejdspladsernes udformning blev ikke ændret under hele undersøgelsen, hvilket gav mulig-
hed for en direkte sammenligning mellem kombinationen af de undersøgte ventilationssy-
stemer. Forskellige sporgasser blev anvendt for at simulere de mest almindelige indendørs
forureningskilder. Der blev valgt gulvbelægning, menneskelig emitterede bioeffluenter og
udåndingsluft. Personers eksponering blev vurderet på grundlag af den koncentration af
gasser, som blev målt i mannequinernes indåndingsluft. Indåndingsluftens temperatur, som
er vigtig for oplevelsen af luftkvalitet, blev ligeledes målt. Den termiske komfort af siddende
individer blev evalueret på basis af varmetabet fra mannequinerne. Yderligere blev forde-
lingen af forurening, temperatur og lufthastighed målt med henblik på at afsløre luftens
strømningsmønstre i rummet. Virkningen af PV systemet blev undersøgt ved forskellige
kombinationer af de personkontrollerede luftstrømme (0, 7 og 15 L/s). Dette svarer til mu-
lige individuelle former for brugen af PV. Ventilationssystemet for rummet som helhed blev
reguleret således, at den totale kølekapacitet af personkontrolleret luft plus luft fra rum ven-
tilationen forblev konstant. Udsugningsluftens temperatur blev fastholdt ved 26°C.

Resultater vedrørende luftkvalitet viste, at i lokaler med opblandingsventilation vil PV altid


være i stand til at beskytte de tilstedeværende personer mod forurening og således forbedre
kvaliteten af indåndingsluften. Princippet for opblandingsventilation indebærer, at typen af
forureningskilde (aktiv eller passiv, lokal eller jævnt fordelt) og dens placering er uden be-
tydning. Kvaliteten af indåndingsluften for personer, der har PV til rådighed, er i stort om-
fang bestemt ved effektiviteten af indblæsningsterminalen, samt retning og omfang af den
personkontrollerede luftstrøm. Indåndingsluftens temperatur var velkorreleret med den i
lokalet jævnt fordelte forureningskoncentration.

I et lokale med fortrængningsventilation viste PV at forbedre kvaliteten af indåndingsluft i


relation til en passiv og jævnt fordelt forurening placeret på gulvet. Dette er tilfældet med
gulvtæppe, PVC og linoleum. Anvendelsen af PV viste sig imidlertid at forøge opblandin-

ix
Resumé

gen af forurening, der befandt sig i nærheden af den personkontrollerede luftstrøm, såsom
udåndingsluft og bioeffluenter. Den personkontrollerede luftstrøm skabte store uensarted-
heder og forskelle i fordelingen af menneskelig forurening tæt ved arbejdspladsen. Længere
væk fra arbejdspladsen forsvandt uregelmæssighederne, men koncentrationen af menneske-
lig forurening blev forøget i sammenligning med tilfældet uden PV. Dette kan om muligt
reducere kvaliteten af indåndingsluften for personer, der ikke har PV til rådighed. Fordelin-
gen af den menneskelige forurening var påvirket af retning og omfang af den personkon-
trollerede luft. Den opadrettede luftstrøm (bordristen) viste sig at medføre en mindre trans-
mission af menneskelig forurening mellem arbejdspladserne end den horisonta-
le/nedadrettete strømning (bevægelige luftfordeler). En personkontrolleret luftstrøm kan
også påvirke fordelingen af forurening frembragt ved en anden arbejdsplads. I så henseende
var indvirkningen af den opadrettede luftstrøm (bordristen) større end indvirkningen af den
horisontale/nedadrettede luftstrømning (bevægelig luftfordeler), mest sandsynligt på grund
af den større strømningshastighed.

Virkningen af PV i kombination med et »underfloor« ventilationssystem med begrænset ka-


stelængde (op til 0.3 m) var sammenlignelig med virkningen af PV i kombination med for-
trængningsventilation. »Underfloor« ventilationssystemet reducerede imidlertid ujævnhe-
der af forurening og temperaturgradienter. Og medførte endvidere, at anvendelse af PV for-
øgede opblandingen af forurening i dens nærhed. Undersøgelserne bekræftede, at en
opadrettet luftstrøm (bordristen) medfører en mindre transmission mellem arbejdspladserne
end den nedadrettede luftstrøm (bevægelige luftfordeler). En forøgelse af systemets kaste-
længde reducerer gradvis effekten af den personkontrollerede luftstrøms retning på overfø-
ring af forurening. Indvirkningen af PV retningen på transmission forsvandt, når kaste-
længden var sammenlignelig med højden af indåndingszonen.

Analysen peger på, at PV, i kombination med et hvilket som helst ventilationssystem, kan
være effektivt til at beskytte personer selv mod meget smitsomme sygdomme, og derfor bli-
ve et alternativ eller supplement til traditionelt anvendte metoder for personbeskyttelse.

Indvirkningen af PV på det termiske indeklima var meget lokalt, selvom endog meget eks-
treme individuelle anvendelsesformer for PV systemet blev undersøgt. Anvendelsen af PV
ved en arbejdsplads havde ikke indvirkning på den termiske komfort for en person ved en
anden arbejdsplads. Den kølende virkning af PV var uafhængig af luftfordelingen frembragt
af ventilationssystemet. Kølingen, som blev frembragt af PV, var ækvivalent med en reduk-
tion af rummets lufttemperatur på 1-2°C, afhængig af den aktuelle kombination af PV ind-
blæsningsterminal, ventilationsprincip og termisk mannequin (arbejdsplads). Muligheden
for, at PV indvirker på luftbevægelser i den nedre opholdszone, var begrænset. I lokaler
med termisk lagdeling (fortrængnings- eller »underfloor« ventilation) viste det sig, at PV
ikke bidrager til at fjerne varme fra den nedre zone. Dette kan medføre en temperaturfor-
øgelse i opholdszonen og herved forøge ventilationssystems energiforbrug. Dog kan en hø-
jere temperatur måske reducere risikoen for træk og betydelige lodrette temperaturforskelle.

Nærværende studie har påvist de kombinationer af PV og rum ventilation, som blev under-
søgt, vil kunne anvendes i praksis. På grund af en fortrinlig virkning på luftkvaliteten og en
lav risiko for termisk ubehag anbefales en kombination af PV og »underfloor« ventilation
med begrænset lodret kastelængde, kontrolleret i overensstemmelse med en strategi for
konstant lufttilførsel. Såfremt kontrol med luftbåren transmission af forurening mellem in-
divider har den højeste prioritet, vil indblæsningsterminaler, der tilfører en opadrettet luft-
strøm, være at foretrække. Udvikling af indblæsningsterminaler med en høj virkningsgrad
og begrænset opblanding af omkringliggende forurening, er anbefalelsesværdig.

x
Nomenclature

Abbreviations
ATD Air terminal device
CAV Constant air volume
CFD Computer fluid dynamics
DR Draught rating
DV Displacement ventilation
MV Mixing ventilation
PV Personalized ventilation
RMP Round movable panel
SBS Sick building syndrome
TV Total-volume ventilation
UFAD Underfloor air distribution
VAV Variable air volume
VDG Vertical desk grill
VOC Volatile organic compound

Symbols
c Contaminant concentration, mg/m3
C Constant
c(-) Normalized contaminant concentration, dimensionless
CRP Contribution ratio of pollution source, dimensionless
ET Manikin-based equivalent temperature, °C
iF Intake fraction, dimensionless
PRE Pollutant removal efficiency, dimensionless
Qt Sensible heat loss, W/m2
RA0 Reproductive number for an infectious disease in indoor environment,
dimensionless
T Air temperature, °C
T(-) Normalized air temperature, dimensionless
Tu Turbulence intensity, %
U Uncertainty
v Mean air velocity, m/s
∆ET Difference between ET with PV and ET without PV; cooling effect, K
εP Personal exposure effectiveness, dimensionless
εV Ventilation effectiveness, dimensionless

Subscripts
0 Without personalized ventilation; reference case
E Exhaust
I Inhaled
PV With personalized ventilation
S Supply

xi
1
1. Introduction

1.1 Background
Numerous field studies (Bluyssen et al., 1996; Mendell, 1993) have documented substantial
rates of dissatisfaction with the indoor environment in many buildings. A recent review of
the scientific literature by a multidisciplinary group of European scientists showed a strong
association between the level of ventilation and comfort (perceived air quality), health (sick
building syndrome (SBS) symptoms, inflammation, infections, asthma, allergy, short-term
sick leave) and productivity (Wargocki et al., 2002). It has been shown that better air quality
increases satisfaction and productivity and decreases health problems. The studies show at
the same time, that meeting today’s standards does not prevent widespread complaints of
poor air quality and frequent building-related symptoms.

Ventilation rates have been given as a required supply of outdoor air per person in order to
dilute human bioeffluents to an acceptable odour level (Fanger, 1998). The fact that many
other sources indoors (new building materials, electronic equipment, etc.) can contribute to
the pollution of air was not considered. Systematic selection of low-polluting materials is
certainly the most efficient way to improve the indoor air quality. However, at present, there
are limits to which the pollution sources can be reduced. Another way to improve the indoor
air quality is to provide larger amounts of outdoor air, but this method poses a risk of
thermal discomfort and may increase energy consumption substantially.

In practice, rooms are used by occupants with different physiological and psychological
responses, clothing, activity, individual preferences to the air temperature and movement,
etc. It is consequently difficult to create an environment that would satisfy everyone when
many people occupy the same space. Various studies have demonstrated that on average, 5-
10% of subjects are always dissatisfied with the thermal environment that is considered
acceptable by the majority of the population. The total-volume ventilation and air-
conditioning, at present the method most used in practice, aims at providing uniform
environmental conditions throughout rooms. The ability of occupants to create and control
their preferred environment is limited.

1.2 Principles of personalized ventilation


Personalized ventilation (PV) provides clean air at each workplace. Its primary aim is to
improve the quality of air inhaled by each occupant, and thus reduce complaints, increase
satisfaction and performance and ensure health. Protecting occupants from inhaling
airborne contaminants, whether they are airborne infectious agents produced by other
occupants or chemicals from building materials, is an important quality of PV systems.

Specifically designed air terminal devices supply clean air direct to the breathing zone (face)
of each occupant. The interaction of personalized airflow with the flow in the human
boundary layer, room airflow and the flow of respiration determines the ability of PV to

1
Chapter 1

ensure high air quality. Low velocity and turbulence intensity of the supply airflow is
required in order to (1) reduce the mixing of the fresh personalized air and the ambient
polluted air and (2) to reduce draught discomfort. However, an airflow that is too weak may
not penetrate the human boundary layer, and it is also susceptible to the impact of buoyancy
forces. The highest air quality is achieved when an occupant inhales air direct from a
potential core region of a personalized air jet. Because of a wide and long concentration core,
the PV terminals developed recently by Bolashikov et al. (2003) are able to provide inhaled
air consisting of nearly 100% clean air from PV at a supply rate of 10 L/s per person.

Fang et al. (1998a, b) confirmed earlier studies and showed that although the odour intensity
of air does not change significantly with temperature and humidity, air is perceived more
acceptable with decreasing temperature and humidity. Because of the positive influence of
low air temperature on the perception of air quality, it is advantageous to supply cool air.
The positive impact of a low air temperature has been confirmed in recent laboratory studies
with human subjects (Kaczmarczyk et al., 2002a, accepted; Zeng et al., 2002; Yang et al., 2003;
Kaczmarczyk, 2003). Cooling of personalized air is not intended to provide space
conditioning.

The presence of air movement in the vicinity of an occupant inevitably affects his or her
thermal sensation. Hence, PV is potentially successful in improving thermal comfort if
individual differences and preferences can be accommodated. The control of thermal
comfort is usually realized by changing velocity, temperature and direction of the supply air
stream. Studies show that PV is able to affect thermal comfort equivalent to decreasing room
air temperature by several degrees (Tsuzuki et al., 1999). This is sufficient for most
occupants. The mechanism of cooling is equivalent to using a desk fan.

Analyses indicate that occupants may have to prioritize between excellent air quality and
preferred thermal comfort. While strong airflow may be desirable in warm climates to
provide cooling, it may at the same time increase the entrainment of polluted room air and
thus decrease the inhaled air quality. In contrast, in a cold environment, sensitive occupants
may prefer not to use PV and thus would not benefit from the high air quality. Kaczmarczyk
et al. (accepted) and Kaczmarczyk (2003) showed that occupants might not perceive the
differences in the air quality between two terminals, which otherwise look similar. He
suggests that thermal comfort is an important parameter for occupant’s preferences.

The provision of individual control is the third advantage of PV systems. Research by


Bauman et al. (1998) suggests that it is more important for workers to have the ability to
control their local environment than it is for them to exactly make a large number of control
adjustments.

The ability to consume less energy in comparison with conventional systems may be the
fourth advantage of PV systems. Although important for application of PV in practice, this
issue has been studied the least. At present, no quantitative proof has been given. The
potential to save energy has been associated with providing smaller volumes of air (less
conditioning), and shutting the PV system off when a workplace is not occupied. As pointed
out by Fanger (2001), the actual breathing requirement of sedentary occupants is as low as
0.1 L/s, while a hundred times more air is typically provided by conventional systems.

In conclusion, personalized ventilation aims to improve perceived air quality and protect
occupants from airborne contaminants, to make each occupant thermally comfortable, to
satisfy his or her individual preferences and to be environmentally responsible in saving

2
Chapter 1

energy. Occupants, in order to avoid draught discomfort, may use PV at small flow rates
and temperatures only a few degrees cooler than the room air temperature. Therefore, total-
volume ventilation and air-conditioning have to be applied in combination with PV in
rooms with high heat and/or pollution loads.

1.3 Design, installation, performance


Individual ventilation and air-conditioning systems have been used in vehicle and airplane
cabins for many years. Air supply nozzles and slots are designed to produce high
momentum jets that promote intensive mixing of the clean air with the surrounding air and
are efficient mainly in improving passengers’ thermal comfort.

In buildings, air terminals incorporated in furniture have been used to deliver conditioned
air close to the occupants in some auditoria and theatres. Applications of PV in offices and
non-industrial working premises have been limited. At present, individually controlled
systems providing an improved level of air quality, referred to also as the task/ambient
conditioning systems, can be grouped in the following three categories:

• Underfloor ventilation systems


• Workplace (furniture) integrated systems
• Air terminals in the breathing zone.

Underfloor ventilation systems


Several researchers investigated the performance of individually adjustable floor-based
terminals named Task Air Modules (TAM), manufactured by Tate Architectural Products.
Air was discharged through four adjustable grills mounted on an access floor panel. A
rotary knob recessed to one grill allowed occupants to control the quantity of supply air.
Fisk et al. (1991) showed that inhaled air quality improved only when the air was directed in
a manner that yielded an upward vertical displacement flow. The thermal comfort
performance of TAM was comparable with other underfloor air ventilation systems (Arens
et al., 1991). Tsuzuki et al. (1999) reported on the ability of TAM to increase the heat loss
from a thermal manikin. The development of another system with the local control of air
velocity and temperature was described by Spoormarker (1990).

Workplace (furniture) integrated systems


Sodec and Craig (1990) reported on desk outlets for office applications designed as linear
slots or round swirl diffusers. The airflow rate was limited to 14 L/s per person in order to
prevent excessive velocities, which could be adjusted in the range from 0 to 0.6 m/s at the
head level of seated occupants. The systems were installed typically in conjunction with an
underfloor air distribution system, which was operated in order to remove the total heat
gain. The improvement of the inhaled air quality was considered but not studied. The
outlets had not seen widespread use due to their cost, and the installations of underfloor
ventilation systems alone prevailed.

Personal Environments Module (PEM) from Johnson Controls has affected a great deal of
research interest. It consists of two adjustable nozzles located at the rear corners of the desk.
From a desktop control unit, the occupant can select his or her own settings for air
temperature, airflow rate and direction. The system is accomplished with a radiant heating
panel at the knee space, task lighting and a background noise generator.

3
Chapter 1

Faulkner et al. (1999) studied the efficiency of PEM to ventilate the breathing zone of seated
occupants. Although the systems could provide re-circulated and filtered air, only the
outdoor air was provided. The system was operated in conjunction with an overhead mixing
ventilation system, which provided additional space cooling but no outside air. The
efficiency was assessed based on the Air Change Effectiveness index (ACE), which was
defined as the ratio of the age of air of the exhaust air and the age of air in the breathing
zone (Etheridge and Sandberg, 1996). The maximum ACE reported for PEM was 1.6, which
indicated 60% improvement in comparison to mixing ventilation. The efficiency was also
studied in regard to the contaminants emitted from the floor covering and the body odours
from other occupants using several point sources of perfluorocarbon tracer gases. The
inhaled air quality was assessed using the Pollutant Removal Efficiency index (PRE), which
is equivalent to the Ventilation Effectiveness index (CEN, 1998). The highest value of PRE
was 1.5 in regard to the body source, and 1.3 in regard to the floor source. The room
distribution of the contaminants was not reported.

In the earlier research by Faulkner et al. (1993), the PEM terminals were operated alone
without an additional system. The supply airflow and temperature were adjusted to balance
the heat load of the office in order to provide a comfortable environment. This was made
possible because of large airflows (up to 40 L/s per person). The inhaled air quality
improved only when the highest rates of outdoor air were directed at the breathing zone.
The conditions were described as unlikely comfortable due to the high velocities in the face.
Measurements throughout the room indicated mixing of indoor air.

Arens et al. (1991) and Bauman et al. (1993) studied the thermal performance of PEM in a
simulated office space. During some tests, a ceiling supply diffuser was used to provide
supplemental space conditioning. The units were shown to be able to maintain close to
comfortable conditions in workstations with different heat load levels. Cho et al. (2001)
measured the air velocity and temperature distribution of PEM installed in combination
with an underfloor air distribution system. The study showed that the temperatures became
more uniform when the supply rate of PEM increased while the airflow from the floor
decreased. The ability of PEM to affect the heat loss from a thermal manikin under various
conditions was studied by Tsuzuki et al. (1999).

Another system available on the market, named Climadesk (manufactured by Mikroklimat


Sweden AB), provides the air from outlets at the front edge of the desk. Two adjustable slots
supply the air horizontally towards the occupant and the third slot, on the top of the desk,
supplies the air upward to the breathing zone. The maximum airflow is about 7 L/s. The
system also provides radiant heating at the thighs. The efficiency of Climadesk to ventilate
the breathing zone of a seated thermal manikin was examined by Faulkner et al. (1999), and
compared to the efficiency of PEM. The maximum values of ACE and PRE in regard to both
a body source and a floor source were 1.9 and 1.6, respectively. The performance of the
Climadesk was a little better than the performance of PEM. The high values were, however,
achieved only when the system supplied outdoor air, and the manikin’s head was located
precisely within the vertical jet of air.

Recently, Faulkner et al. (2002) investigated experimentally the effectiveness of an air supply
nozzle located underneath the front edge of the desk. The supply airflow rates ranged from
3.5 to 6.5 L/s. The measured values of ACE in the breathing zone ranged from 1.4 to 2.7,
which converted to PRE (using the correlation between ACE and PRE for a floor source
presented in Faulkner et al., 1999) correspond to 1.2 and 2. The efficiency was higher than

4
Chapter 1

typically reported for previously tested outlets (Faulkner et al., 1999) or displacement
ventilation systems. The system was tested in a room with mixing ventilation.

Loomans (1998) proposed a desk displacement ventilation system. The supply terminal was
situated below the desktop (against the back of the desk) and supplied air at a low velocity,
0.1 to 0.2 m/s, towards the occupant. The concept was tested experimentally and
numerically. The inhaled air quality reportedly did not improve in comparison to rooms
with traditional displacement ventilation. Izuhara et al. (2002) tested a similar concept. A
fan-equipped partition was used to deliver clean air from the floor level under the desktop.
A traditional displacement ventilation system was installed in the room. The measurements
using the age of air concept did not show large differences between the displacement
ventilation alone and in combination with the partition.

Studies on PV and task/ambient conditioning systems reported in the literature recently


include a ventilation tower system (Hiwatashi et al., 2000) and a partition integrated fan-coil
unit (Chiang et al., 2002). An improvement of air quality and/or thermal comfort for
occupants was indicated; however, the studies did not reveal exceptional characteristics or
outstanding performance of the designs tested.

Recently, Melikov et al. (2002) investigated five different designs of supply air terminal
devices. The designs are schematically shown in Figure 1.1. They comprised: Movable Panel
placed above occupant’s head, Computer Monitor Panel located above the monitor, two
desk grills supplying air vertically upward and horizontally towards the torso from the front
edge of the desk, and pair of PEM terminals tested previously by Faulkner et al. (1999). The
terminals were adjusted in order to provide clean air directly to the breathing zone. The idea
of the movable panel was that due to its construction it is possible to change the location of
the terminal in relation to the occupant.

Figure 1.1. Air terminal devices studied by Melikov et al. (2002): movable panel (MP), computer
monitor panel (CMP), vertical desk grill (VDG), horizontal desk grill (HDG) and personal
environments module (PEM).

The terminals were compared in terms of the inhaled air quality and thermal comfort of a
seated occupant, using a breathing thermal manikin in a climate chamber. The room
distribution principle was upward piston flow with a mean velocity of less than 0.06 m/s.
An isothermal and 6 K lower than the ambient chamber air was supplied from the terminals
at an airflow rate ranging from 3 to 23 L/s. A new index, Personal Exposure Effectiveness,
was proposed. The efficiency of personalized ventilation is expressed as the portion of clean

5
Chapter 1

personalized air in inhalation. Furthermore, the temperature of inhaled air was measured.
Results showed that the performance of PV in terms of both the indices depends to a large
extent on the supply air terminal type and the airflow rate. Increasing the airflow rate
through the PV increased the amount of personalized air inhaled. The highest personal
exposure effectiveness was about 0.7 (PRE of 3.3) measured with the panel above the
computer screen (CMP). The supply airflow was 23 L/s and isothermal. The best performing
outlet for both an isothermal and a non-isothermal air supply was the vertical desk grill
(VDG), which provided a personal exposure effectiveness of about 0.5 (PRE of 2.0) already at
an airflow rate of 10 L/s. The next best outlet was the movable panel (MP), which for
isothermal conditions performed similarly to the VDG. The study also reports on the portion
of exhaled air that was re-inhaled. It was generally low and did not exceed 1% for any of the
conditions or terminals tested. The cooling ability of the terminals was reported as well.

Kaczmarczyk et al. (2002a, b) modified the Movable Panel in order to provide better
flexibility and appearance of the system (later referred as the 2nd generation MP). The human
response to the terminal was examined (see below). The personal exposure effectiveness
(measured additionally) was 0.3, which is, however, rather mediocre. This, together with
somewhat encouraging results from the experiments with human subjects, initiated a
development of a new terminal (Bolashikov et al., 2003), named Round Movable Panel
(RMP). It was made in a circular shape and fitted with a flow straightener in order to
provide low turbulent flow with a long concentration core. The RMP was mounted on a
movable arm-duct attached to the desk, which allowed for its positioning in respect to the
occupant. Physical measurements revealed that for the typical positioning, as used by
occupants, the inhaled air consisted of 100% of clean personalized air at a supply rate of 15
L/s. At lower rates, the performance was influenced by the temperature difference between
the supply air and ambient air.

Air terminals in the breathing zone


The effort to reduce the mixing of clean personalized air and ambient air led to a
development of terminal devices positioned direct in the breathing zone of a person. Zuo et
al. (2002) studied the concept of a facial air supply outlet. Nozzles of different shapes and
sizes were placed on the chest of a manikin and provided with air at a rate of up to 2 L/s. At
the highest rate, the ratio of the personalized air in inhalation was calculated to be 0.61. The
results are compromised by the fact that the manikin was not heated. It is known that the
inhaled air quality depends on the complex interaction of airflows around a human body, of
which the free convection along the body is one of the most important.

Bolashikov et al. (2003) developed an air terminal incorporated in a commercially available


set of headphones. The microphone piece was replaced with a small rectangular nozzle,
providing clean air of up to 0.5 L/s from a short distance direct to the mouth/nose of a
person. The design was tested by means of both physical measurements with a breathing
thermal manikin and experiments with human subjects (see below). The physical
measurements revealed that the inhaled air consisted of up to 80% of clean air from the PV.
The ability of the Headset to affect the heat loss from a thermal manikin was very small.

Human response to personalized ventilation


Knowledge about the human response to PV is limited. Kaczmarczyk et al. (2002a, b,
accepted), Zeng (2002) and Kaczmarczyk (2003) examined the response of 60 human subjects
to the 2nd generation Movable Panel (MP) and mixing ventilation at several combinations of
room air temperature and PV air temperature. A series of 4-hour experiments was carried
out in a controlled laboratory environment. They showed that MP providing cool outdoor

6
Chapter 1

air was able to improve the perceived air quality and decrease the intensity of some SBS
symptoms compared to mixing ventilation. The acceptability of the thermal environment
with PV compared to the situation without PV was improved at the higher temperature in
the office. Improved self-estimated performance was indicated.

Most recently, Kaczmarczyk (2003) and Kaczmarczyk et al. (2004) examined perceived air
quality, thermal sensation and individual preferences with five different air terminal devices
for PV and mixing ventilation. The devices used were the Round Movable Panel (RMP), the
Headset, the 2nd generation MP, the Vertical Desk Grill (VDG) combined with the Horizontal
Desk Grill (HDG), and the Round Movable Panel combined with the Horizontal Desk Grill.
The subjects experienced each system in 25-minute sessions under different combinations of
the pollution level and temperature of ambient air. Results showed that all the terminals
improved the perceived air quality compared to the office (mixing ventilation). The system
with the highest rating in terms of the perceived air quality and thermal conditions was the
RMP. It was also the system most frequently selected by subjects. The subjective evaluation
of the MP was fairly similar to the RMP, despite the differences in their ventilation
performance identified by objective physical measurements. The combination of VDG and
HDG caused unpleasant cooling of the pelvis, legs and chest, which was due to HDG. The
analyses showed that most subjects used the VDG to achieve high air quality, and closed the
HDG. The least rated terminal was the Headset. The subjects reported an unpleasant
sensation due to very localized and high velocity, impractical and uncomfortable design (air
was supplied to the headset through tubing), and lack of cooling for larger body areas at
elevated temperatures.

At present, the movable panels, namely the RMP, and front-desk-edge mounted grills have
the highest potential to ensure excellent air quality and preferred thermal environment for
occupants. The performance of PV in regard to the distribution of different contaminants has
been studied only recently (Cermak and Melikov, 2003; Faulkner et al., 1999).

1.4 Contaminants indoors


All people emit a complex mixture of effluents, which produce an unpleasant odour in
sufficient concentration. The odour levels are typically controlled by ventilation to a level
that is acceptable by most occupants. The percentage dissatisfied with air polluted by
human bioeffluents as a function of the ventilation rate per person is available (CEN, 1998).
Although the bioeffluents are not the strongest pollutant in today’s buildings, they are still
important because they will ultimately be present in buildings after other contaminants have
been removed.

Carbon dioxide (CO2) produced in the human lung proportionally to the metabolic rate is a
good indicator of human bioeffluents. CO2 is not toxic, except at high concentrations.
Exhaled air contains about 3.6% CO2, which is, however, diluted in indoor air by ventilation.
The first effects are noticeable at a concentration of about 1% (McIntyre, 1980). This is rather
high and reached typically only in crowded and under-ventilated spaces. The symptoms
include increase in the depth of breathing, frequency of breathing and headache, which may
reduce performance.

Environmental tobacco smoke is an unprecedented source of odour and irritation. It has


been shown to increase the risk of a variety of diseases and no safe levels of exposure can be
recommended. Despite its adverse health effect, tobacco smoke continues to be one of the

7
Chapter 1

most important contaminants indoors, especially in underdeveloped parts of the world. Its
emission is a combination of the smoke exhaled by the smoker and the smoke released
directly from the burning cigarette.

Air exhaled by people is the vehicle for release of respiratory infectious agents. Both viruses,
which range in size from 0.003 to 0.06 µm, and bacteria, which mostly range between 0.4 and
0.5 µm, do not occur alone but in colonies or attached to other particles. The agents can be
dispersed from the respiratory tract during sneezing, coughing or talking.

Figure 1.2. Predicted total respiratory depositions at three levels of exercise based on the International
Commission on Radiological Protection deposition model. Average data for males and females.
Reprinted from Hinds (1999).

Most airborne viral and bacterial aerosols originate from human-produced droplet nuclei.
Very little information is, however, available on the size of these droplets. The distribution
reported in the literature differs according to the measurement techniques applied. Duguid
(1945), who measured stain marks found on slides exposed to exhaled air, reported that
most droplets were between 4 and 8 µm in diameter. Fairchild and Stamper (1987)
concluded, using an optical particle counter, that a great majority of particles are less than
0.3 µm and a few greater than 1 µm. Papineni and Rosenthal (1997) demonstrated the
existence of droplets in the exhaled breath using an optical particle counter and
measurements of dried droplets collected upon electron microscopy grids. The droplet size
ranged from the lower limits of detection of the methods used (0.3 µm with the optical
counter) to approx. 8 µm. After expulsion, the large droplets either settle out of the air or
they evaporate to droplet nuclei that approach the size of the individual agent. Brosseau et
al. (1994) reported, based on a literature review, that the diameter of droplet nuclei ranges
from 0.5 to 5 µm. ASHRAE Systems and Equipment Handbook (2000) states that the droplet
nuclei average about 3 µm in diameter.

The exhaled air may also contain particles that were previously inhaled and did not deposit
in the respiratory tract. The respiratory deposition for a wide range of particle sizes
presented in Figure 1.2 indicates that the most respirable particles range from 0.1 to 1 µm.

8
Chapter 1

Infectious particles behave physically in the same way as any other aerosol-containing
particles with similar size, density, electrostatic charge, etc. Particles less than 0.1 µm in
diameter behave similarly to gas molecules. They travel with Brownian movement and with
no predictable or measuring settling velocity. Particles from 0.1 to 1 µm have settling
velocities that can be calculated but that are very low. Although particles in the 1 to 10 µm
range settle in still air, normal air currents may still keep them in suspension for appreciable
periods. Thus airborne infectious agents can be transported by airflow from person to
person. The successful transmission of an infection depends on susceptibility of the
individuals (immunity), duration of exposure, concentration of agent, virulence of agent, etc.

Volatile Organic Compounds (VOCs) have been associated with poor air quality, eye and
airway irritation and consequently the prevalence of SBS symptoms (Wargocki et al., 1999;
Pejtersen et al., 2001). Building products, which represent the largest surfaces indoors, are
considered the major VOC sources (Wolkoff, 1995). Most studies associated the building-
related complaints with the presence of carpeting, PVC and linoleum (Mendell, 1993;
Jaakkola at al., 1999, 2000; Wolkoff et al., 1995). Despite availability of low-polluting
materials, such as polyolefin, carpets are still being used frequently in many buildings.

Recently, some personal computers (PC) have been identified as a strong pollution source
having a negative effect on perceived air quality and productivity in offices (Bakó-Biró et al.,
2004). Sensory evaluation showed that the classical cathode-ray tube displays were the
major polluting elements. Polluting load of the flat Thin Film Transistor (TFT) displays as
well as computer towers was small (Wargocki et al., 2003). The significance of PCs as
contaminant sources is expected to decrease due to an increase of the market share of the
TFT displays. Other office equipment such as printers and copy machines producing ozone
are typically fitted with active charcoal filters, which reduce the ozone emission. Besides,
they are often placed in sparsely occupied areas (e.g. corridors) where the exposure of
occupants is limited.

The exposure of occupants to a contaminant depends on the distribution of the contaminant


in a room. The distribution is influenced by the principle of ventilation, type of a
contaminant source and its location in respect to the occupant, airflow generated by the
activity of occupants, distribution of heat sources (thermal plumes), weather conditions, etc.

1.5 Room air distribution


Total-volume ventilation and air-conditioning of rooms is at present the method most used
in practice. Mixing and displacement room air distribution are the main principles applied.
Underfloor air distribution, which combines the characteristics of mixing and displacement
distributions, has become popular for ventilation of offices in recent yeas.

1.5.1 Mixing ventilation


Mixing air distribution aims at creating relatively uniform air velocity, temperature,
humidity, and air quality conditions in the occupied zone. Air quality is maintained by
dilution of the released contaminants. Conditioned air is normally supplied from air
terminal devices at relatively high velocities, much greater than those acceptable by building
occupants. Supply air temperature may be above, below or equal to the air temperature in
the occupied zone. The diffuser jets mix with the ambient room air by entrainment and
reduce the air velocity and equalize the air temperature. With ceiling-based devices, a region
of discomfort is contained above the head level and does not typically affect the occupants.

9
Chapter 1

However, the quality of air inhaled by occupants is necessarily lowered, allowing the supply
air to mix with the contaminants that collect near the ceiling.

Studies show that ideal mixing may not be and is often not achieved. Fisk et al. (1997)
provided evidence of short-cutting of air between supply diffusers and exhaust grills located
on the ceiling when the supply air was heated, especially when ventilation rates were low.
The short-cut did not occur when the supply air was cooled. Heiselberg (1996) showed that
the supply airflow rate, location of the return opening, location of the contaminant source
and density of the contaminant influences the contaminant distribution. Large differences in
distribution were found especially when the airflow rate was low. On the other hand, the
activity of occupants (walking, opening and closing of doors) contributes to mixing and
diminishes the large differences in air quality that are often found in the test rooms.

CEN Report 1752 (1998) acknowledges an inhomogeneity of the air quality in mixing-
ventilated rooms. The typical values of ventilation effectiveness range from 0.4 to 1. The
lower values are achieved when the temperature of supply air is higher than the
temperature of air in the breathing zone. The low values of ventilation effectiveness are
associated with heating. In cooling application, the ventilation effectiveness typically ranges
from 0.9 to 1.

1.5.2 Displacement ventilation


Nielsen (1993) and most recently REHVA (2002) provide comprehensive overviews of
displacement ventilation (DV). Ventilation air is introduced direct to the occupied zone at a
temperate slightly (3-6°C) below the room temperature by floor- or wall-mounted diffusers.
Due to the gravity forces, it spreads along the floor in an almost horizontal layer. The
thermal plumes generated by warm surfaces (people, equipment, etc.) entrain and transport
the air as well as heat and contaminants from the lower levels of the space upward, where
they are exhausted at or close to the ceiling. The fraction of air that is not exhausted (the
exhaust rate is usually smaller than the flow generated by the plumes) is forced to flow
downwards and mix with the ambient air. A level that is said to separate the clean and
polluted parts of the room is called a neutral or stratification height.

Displacement air distribution has the ability to provide occupants with better air quality, as
compared with traditional mixing systems, especially when the contaminant sources are also
heat sources (Brohus and Nielsen, 1996). Energy can be used more efficiently, because air is
exhausted at temperatures that are several degrees above the temperature in the occupied
zone. However, vertical air temperature differences always exist in displacement-ventilated
rooms, with low air temperatures and high air velocities often near the floor. Therefore, if
not well designed, the risk of local discomfort due to draught (Pitchurov et al., 2002) and
vertical air temperature difference is high (Melikov and Nielsen, 1989).

Contaminant distribution
The undisturbed flow pattern gives a gradient in both temperature and contamination
within the room. The gradients are not necessarily of the same form. The characteristic two-
zonal contaminant distribution is generated when the contaminant sources are associated
with heat sources (Brohus and Nielsen, 1996). This is the case of e.g. electronic equipment or
human bioeffluents emitted from a still person. The interface layer between the lower clean
and upper polluted zone is formed where the net flow rate of plumes equals the supply
airflow rate. The thickness of the layer is typically about 0.5 m (Etheridge and Sandberg,
1996). The amount of air transported in the convection flows, and the height to which the

10
Chapter 1

plume rises, depend on the shape, surface temperature and distribution of the heat sources.
Furthermore, the high to which the plume rises is strongly influenced by the temperature
gradient in a room. If the convection current from the contaminant source, in a room with
several heat sources, is not the warmest, the contaminant may settle in a layer where the
concentration locally exceeds the exhaust concentration (Skistad, 1994). Data on the volume
flux in the plumes above different sources can be found e.g. in Nielsen (1993), Mundt (1995)
or Aksenov et al. (1998).

In a calm environment a free convection flow exists around warm or cold surfaces. Around
people, the free convection flow is able to a great extent to entrain and transport air from the
lower part of the space to the breathing zone. Brohus and Nielsen (1994, 1996) proposed a
quantity named the effectiveness of entrainment in the human boundary layer, which
expresses the ability of the free convection flow to supply (fresh) air from the floor area to
the breathing zone. An almost linear relationship between the effectiveness and the ratio of
the stratification height to the breathing zone height was identified. However, the inhaled
air quality in displacement ventilated rooms may not be high when contaminants are
passive, i.e. without any significant initial momentum or buoyancy, and/or positioned close
to a person (Brohus and Nielsen, 1995, 1996). Murakami et al. (1998) introduced a modified
effective entrainment ratio, which takes into account the ability of the rising stream to carry
not only clean air but also contaminants, which decrease the inhaled air quality. Hayashi et
al. (2002) performed CFD analyses showing the spatial distribution of the portion of air to be
inhaled by a standing, sitting and sleeping occupant.

Unless the source is located on the floor or close to an occupant, the values of ventilation
effectiveness typically exceed 1. Examples of ventilation effectiveness in the breathing zone
as reported in CEN Report 1752 (1998) range from 1.2 to 1.4, assuming that the contaminant
sources are distributed uniformly in the space. The results of various full-scale
measurements reported in the literature are summarized in Brohus and Nielsen (1996). The
values of ventilation effectiveness ranged from 1 to 8 (referred to as the personal exposure
index).

Activity of occupants
The activity of occupants in a displacement-ventilated room is generally detrimental to the
inhaled air quality. Movements of a person can disturb the boundary layer around his/her
body, which prevents the breathing zone from being supplied with usually clean air from
the lower space when the person is still. A person moving/walking in the room may also
disturb the overall temperature and contaminant distribution and affects thus the inhaled air
quality of other people in the room indirectly.

Hyldgaard (1994) and Brohus and Nielsen (1995) studied the concentration of contaminants
inhaled by a thermal manikin exposed to uniform horizontal flow of different velocities in a
wind channel. The effect of movements of the manikin was assumed to be equivalent to the
impact of the uniform velocity field. They showed that the boundary layer and hence the
inhaled air quality was affected considerably already at a velocity of 0.1 m/s. Mattsson
(1999) and Bjørn et al. (1997) carried out a study with a person simulator moving
continuously back and forth in a displacement-ventilated room. The inhaled air quality
decreased at around 0.2 m/s, at which speed the convection flow reportedly seemed to be
deflected away from the breathing zone. All the studies generally agree that the air quality
in the breathing zone of a fast walking person (> 1.0 m/s) can be considered the same as in
the ambient air.

11
Chapter 1

The ambient air quality in the occupied zone depends strongly on the physical activity of the
people in the room. Brohus and Nielsen (1994) measured the concentration profiles with
four persons, two of whom were either seated or walked normally around the room. They
showed that although the concentration profile changed between the two conditions, the
moving people were not able to destroy the stratification completely. Mattsson (1999)
showed that a rather high activity, such as a sports activity, is needed to completely abolish
the displacement effect. Physical measurements with the back and forth moving human
simulator reported by Mattsson (1999) and Bjørn et al. (1997) showed that already at a speed
of 1.0 m/s the distribution was close to the completely mixed situation.

Exhaled air
The distribution of exhaled air deserves special attention because of its importance with
respect to the transmission of infectious agents, and in situations with passive smoking. Air
is exhaled with positive buoyancy and initial momentum. It typically penetrates the free
convection boundary layer around the body and becomes free of it. Observation shows that
both the buoyancy and momentum are diffused quickly after the exhalation. In a calm
environment the exhaled air may stratify in the breathing zone height. If it does so, the local
concentration may exceed several times the concentration around the person at the same
height. Different authors disagree about the impact of a breathing opening. Bjørn et al.
(1997) observed stratification of air exhaled through the mouth. Exhalation through the nose
did not stratify and the contaminant distribution was similar to the case when the
contaminant was released in the plume above the manikin. To the contrary, exhalation
through the nose reportedly stratified in experiments of Hyldgaard (1994). Bjørn (2002)
showed that the pulmonary ventilation rate is more important for the flow pattern in front
of a person than the exhaled air temperature. According to Bjørn (2002), the stratification is
affected by the steepness of the vertical temperature gradient in the immediate surroundings
of the respiration zone. The critical limit for the stratification to develop is approx. 0.5 °C/m.
Bjørn et al. (1997) showed that movement of a manikin in the room at a very low speed (0.2
m/s) dissolved the stratification layer of exhaled air.

Bjørn and Nielsen (1996) studied personal exposure to air exhaled by another person using
two breathing thermal manikins standing in a displacement-ventilated room. They showed
that the inhaled air concentration was significantly greater than in the exhaust when the
manikins exhaled directly towards each other. As the distance between the manikins
increased, the exposure decreased. The concentrations inhaled were comparable to the
exhaust concentration when the distance exceeded 1.2 m for exhalation though the mouth,
and 0.8 m for exhalation through the nose. When exhalation was directed towards the back
of the manikin, larger exposures did not occur. A CFD simulation by Bjørn and Nielsen
(1998) showed that the personal exposure was very sensitive to variations in the convective
heat output of both the exposed person and the exhaling person, and in the cross-sectional
exhalation area (mouth) and the pulmonary ventilation rate of the exhaling person.

Particles
Only the distribution of contaminants that follow air currents, such as gases, vapours and
small particles, has been considered so far. The ability of DV to displace the contaminants
that do not follow the currents, namely larger particles, to the upper part of the room and
create a fairly clean lower zone may be limited. Mattsson (2002) studied the vertical
distribution of particles generated through office-like activity of people ranging from 0.3 to >
25 µm. The distribution was measured in a displacement-ventilated room at moderate to
high ventilation rates. He showed that the displacement effect started to decline for particles
in the range 5-10 µm. Slightly negative concentration gradients were observed for particles >

12
Chapter 1

10 µm at the lowest ventilation rate (1.5 air changes per hour), suggesting a significant
influence of gravity. Mundt (2001) obtained similar results.

Thermal environment
A vertical temperature gradient will rise in a room due to the vertical flow of heat to the
ceiling region. The gradient is always positive, increasing temperature up to the ceiling. Its
magnitude and shape depend strongly on the geometrical extension, surface temperature
and vertical location of the heat sources, supply airflow rate (Mundt, 1995) and the
emissivity of the surfaces in a room (Nielsen, 1995; Li et al., 1993). Temperature gradients for
different rooms and heat source arrangements are presented elsewhere. It is a general
experience that the vertical temperature gradients are identical at any location in the room
outside areas with thermal plumes. Nielsen (1995) summarized five temperature
distribution models with varying levels of complexity. The temperature gradient is often
assumed to vary linearly with the height from a minimum temperature near the floor to a
maximum temperature near the ceiling. Mundt (1995) showed that the linear model is a
good approximation in rooms of moderate heights.

The thermal stratification in the room air is much less influenced by physical activity than
the contaminant stratification, apparently due to the accumulation of heat in materials
(Mattsson, 1999; Brohus and Nielsen, 1994). The stratification is usually quite stable, even if
people are moving around in a room. A stronger gradient is less sensitive to the
disturbances. Even if the room air is temporarily mixed, the stratification is re-established
after cessation of the activity.

Air distribution in the vicinity of terminals


In order to avoid thermal discomfort it is necessary to be aware of the adjacent zone close to
the air terminal device, so-called near zone. Several definitions are nowadays used in
practice. Originally, the near zone was considered as a zone around the outlet within which
the mean velocity is higher than 0.2 m/s. In some cases the near zone was restricted only for
heights 0.1 m above the floor level. Melikov et al. (1990) proposed a more reasonable
definition of the near zone based on the percentage dissatisfied due to draught (Fanger et al.,
1988; ISO, 1994). Melikov and Langkilde (1990) showed that the zone for 15% dissatisfied
penetrates almost twice as deep into the room as the zone defined by the mean air velocity
of 0.2 m/s. The size and the shape of the near zone are a characteristic of each air terminal.
The information is usually available from a manufacturer.

Practical considerations
There are several disadvantages associated with the DV concept for offices.
• The lowest permissible supply air temperatures restrict the cooling capacity of DV to 30-
60 W/m2, depending on the air terminal design. In contrast, overhead mixing ventilation
systems can remove loads of up to 100 W/m2 comfortably.
• Displacement ventilation is not suitable for heating.
• Large airflow rates of clean air are required in order to maintain a high level of
stratification, which is necessary to ensure a high quality of inhaled air.
• There are large space requirements in respect to the near zone. The areas can neither be
occupied nor furnished.
• The principle ensures horizontal uniformity and limits individual control over the
environment.
• The principle may be inefficient in regard to passive pollution sources.

13
Chapter 1

1.5.3 Underfloor air distribution


UFAD systems use an open space between the structural slab and the underside of a raised
floor to deliver ventilation air direct to the occupied zone. The air is delivered to the space
through floor diffusers or grills and returned at or close to the ceiling. An UFAD system may
offer improved air quality levels in the occupied zone due to the stratification of
contaminants and reduced energy. Moreover, UFAD systems may better address the
individual differences in regard to thermal comfort as well as provide flexibility of
relocation and reconfiguration of workplaces. Loudermilk (2003) states, however, that cases
that involve justification of a raised floor system solely based on the integration of an
underfloor air delivery system are few. Instead, economic justification (in contrast to
overhead mixing systems) is usually achieved upon the flexibility the platform offers to the
relocation of incorporated power, voice and data services for the space.

Several factors affect the design and thus the typical air distribution pattern of underfloor
ventilation systems. First, there is an increased possibility of extensive draught as well as
vertical temperature difference due to the close proximity of supply outlets to occupants.
The requirement for a small near zone facilitates high induction air terminals, which
discharge air most efficiently by means of turbulent jets. The jets are often designed to
develop in the vertical direction in order to reduce the extent of the near zone. The thermal
comfort issues practically limit the supply air temperatures to 15-18°C, and the quality of
supply air up to 50 L/s per diffuser. This increases the number of diffusers that need to be
used as compared to the number of ceiling-based diffusers used in room with overhead
mixing ventilation. Consequently, numerous diffusers providing turbulent jets reportedly
create mixing conditions in the lower part of a room (Loudermilk, 1999).

The height above the floor where supply air stream velocity decreases to 0.25 m/s is referred
to as the throw height or a maximum penetration height of supply air jets. It is a common
assumption that the mixing effect is then minimized. Above the mixing zone, air is entrained
into the heat sources and drawn upward in the form of thermal plumes, as in the case of
displacement ventilation. The plumes then transport heat and contaminants to the upper
space of the room where a mixed zone is typically established. A relation between the throw
height and the stratification height is crucial for the air quality performance of UFAD
systems. Figure 1.3 or illustrates two scenarios that may happen. Assuming there is a step
change of concentration from the supply air level to the exhaust level, and the interface
between the lower clean and the upper polluting zone is higher than the breathing zone
height, the inhaled air quality would equal the supply air quality (Figure 1.3, left). The
contaminants are assumed to be associated with heat sources (e.g. human bioeffluents). If
the interface height is below the breathing zone height, but still higher than the throw, the
inhaled air quality could be predicted using the model of Brohus and Nielsen (1996) for
displacement ventilation.

Yamanaka et al. (2002) proposed a model to calculate the vertical distribution of


contaminants, when the maximum throw is greater than the height of the interface. The
model was verified experimentally in a scaled room. Providing that the supply airflow does
not affect the thermal plumes around occupants (i.e. they can transport air upward), it
interacts with the interface and forces the contaminants from the upper polluted zone to
flow downward (Figure 1.3, right). The model allows for predicting the contaminant
concentration in the lower occupied zone and the thickness of the interface layer. Occupants
were considered the only sources of heat and contaminants.

14
Chapter 1

Figure 1.3. Air distribution models for underfloor air distribution.

The vertical throw and the extent of the near zone depend on the characteristics of a diffuser
(geometry), the supply air volume and the temperature difference between supply air and
room air. Several types of floor diffusers are recognized. Swirl diffusers supply air with
high-turbulence flow and a large induction effect; they are the diffusers most often available
commercially. Constant air velocity (active) diffusers are designed for VAV operation. An
internal automatic damper maintains a constant discharge velocity, even at reduced supply
air volumes. The throw is relatively constant as the discharge area is throttled. The diffuser
may also consist of a slotted square floor grill supplying the air in a jet-type pattern.
Displacement floor outlets generate low-turbulence, radial horizontal flow for displacement
ventilation. For many years, linear floor grills have been used, particularly in computer
room applications. Air is supplied in a jet-type planar sheet making them suitable for
placement in perimeter zones adjacent to exterior windows.

The selection of floor diffusers should be based on the depth of the diffuser’s mixing zone
and the radius of the near zone. Bauman and Webster (2001) recommend designing the floor
diffuser systems to allow mixing only in the occupied zone, i.e. up to 1.2-1.8 m. Loudermilk
(2003) argues that the mixing must be limited to just below the respiration level in order to
transmit the exhaled air to the exhaust and thus prevent cross-respiration to other
occupants. The mixing zone depth is also critical for creating the upper level stratification
necessary to efficiently isolate space convective loads (Loudermilk, 1999). Consideration
must be taken to ensure that outlets are not located so close to stationary occupants because
uncomfortable conditions around the outlets are likely to bother them.

There are not many studies on the thermal performance of UFAD systems either. A
similarity of underfloor ventilation with displacement ventilation is often anticipated. No
particular attention was paid to the impact of the vertical throw on the thermal
environment. Webster et al. (2002a, b) studied the thermal stratification with variable area
and swirl diffusers in a test room. They showed that as the total room airflow increases at a
constant heat load, room air stratification decreases and vice versa, particularly for the swirl
diffusers. The reason might be the change of the diffusers throw. The temperature near the
floor remained relatively constant unless the supply air temperature changed. Variations of
the supply air temperature increased or decreased the temperature profiles, but it did not
affect their shape. The variable area diffusers were less sensitive to diffuser flow rate,
because they maintained a constant throw estimated at 1.8 to 2 m. Bauman and Daly (2003)
state that the normalized temperature near the floor increases from 0.5 with displacement
ventilation to 0.7 with UFAD because of mixing in the occupied zone.

15
Chapter 1

Despite the lack of deeper understanding of the room air distribution with UFAD systems,
mixing and displacement air distributions are the extreme cases. The literature (e.g. Arens et
al., 1991; Bauman et al., 1991; Fisk et al., 1991; Loudermilk, 1999; REHVA, 2003; Bauman and
Daly, 2003) indicates that the airflow pattern of UFAD resembles displacement ventilation
when a discharged strategy with a short vertical throw is employed. On the other hand, high
airflow rates supplied vertically may reach the ceiling and produce mixing, which is
comparable to overhead systems. The most recent overview of an underfloor air distribution
(UFAD) can be found in Bauman and Daly (2003), who address also many practical issues
such as the design of the supply plenums, UFAD equipment, energy use, etc.

1.5.4 Control strategies


There are two basic strategies to maintain a comfortable environment in response to
changing heat load by supply air. A Constant Air Volume (CAV) system changes the supply
air temperature in response to the space load, while maintaining a constant airflow. The
indoor air quality remains the same unless the contaminant emission varies.

With a Variable Air Volume (VAV) system, a terminal unit at the zone varies the quantity of
supply air to the space. The supply air temperature is held relatively constant. Supply air
temperature must always be low enough to meet the cooling load in the most demanding
zone, and to maintain appropriate humidity. The ability to reduce the supply air quantity
allows for energy savings on the one hand, and on the other it reduces the air changes per
hour and thus decreases the indoor air quality. The VAV strategy can also be operated on
the basis of contaminant concentration, such as CO2. The airflow rate is then varied in order
to maintain air quality, while changes in temperature maintain the thermal environment.

1.6 Interaction of airflows


Airflows of a different nature exist in the vicinity of an occupant. The free convection flow
around a human body, the flow of respiration and the flows generated by PV, and the total-
volume ventilation systems are the most important. The complex interaction of these
airflows determines both the inhaled air quality and thermal comfort. Section 1.5 presented
the airflow patterns with the most common ventilation principles.

Transient flow of respiration


The respiration depends primarily on the activity level and the body weight. At light work a
seated person of an average size has a respiration frequency of 10 per minute. Each cycle of
the breathing function consists of inhalation, exhalation and pause. The pulmonary
ventilation (airflow rate) is 0.6 L per inhalation and the typical breathing pattern is through
the nose. The exhaled air has a temperature of approximately 34°C (influenced by room air
temperature) and a relative humidity close to 100% (Höppe, 1981).

The inhaled air quality is affected mostly by the flow of inhalation. The flow has only a small
impact on the airflow around the human body due to the rapid velocity decay near the
opening. Hyldgaard (1994) calculated that assuming the air is taken from half a sphere, the
velocity already at a distance of 0.05 m from the mouth decreases to 0.015 m/s. Hence, the
impact of inhalation through the mouth or through the nose on the inhaled air quality is
negligible.

The flow of exhalation affects the air movements around the human body much stronger
than the flow of inhalation. Hyldgaard (1994) performed velocity measurements along the

16
Chapter 1

axis of the exhaled air from the nose air jet with the head in upright position. He reported
that the mean velocity of the exhaled air from each nostril is 1.85 m/s. The direction of the
exhalation is approximately 45° below the horizon. The two nostrils create two independent
jets 30° apart which do not collapse but diffuse in the room. Section 1.5.2 described the
distribution of exhaled air in a room with displacement ventilation.

Free convection flow around a human body


Upward free convection air movement around the human body exists due to differences
between the temperature of the body surface and the room air temperature. The airflow is
slow and laminar with a thin boundary layer at the lower parts of the body and becomes
faster and turbulent with a thick boundary layer at the height of the head. The velocity
profile is similar to a wall jet. Homma and Yakiyama (1988) measured a maximum velocity
and the boundary layer thickness at the ankle and beside the head of a standing naked
subject. The maximum velocity and the boundary layer thickness were 0.1 m/s and 8 mm
for the ankle level and 0.25 m/s and 150-200 for the head level, respectively. The room air
temperature was 20°C. Melikov and Zhou (1996) measured the velocity and temperature
distribution at the neck of a seated and clothed thermal manikin using a highly accurate
Laser Doppler Anemometer. The maximum velocity was 0.18 m/s at a distance of 5-8 mm
from the surface. The velocity boundary layer thickness was 80-90 mm. The thickness of the
temperature boundary layer was 30-35 mm.

The air movement caused by the free convection flow is important for people’s thermal
comfort and air quality. As described earlier in Section 1.5.2, the flow in the human
boundary layer is able to entrain and transport air (polluted or clean) from the lower levels
to the breathing zone, from where it is inhaled. Furthermore, the air is heated, humidified
and polluted by bioeffluents generated by the body and it is distributed to the room in the
form of a thermal plume.

Personalized airflow
Airflow generated by personalized ventilation can be very different. There is primary
airflow supplied from an air terminal device, but also secondary airflows generated by
entrainment in a confined workstation space. As stated earlier, the personalized airflow
should be of low velocity and low turbulence intensity in order to reduce the mixing of clean
and polluted air. The primary air stream should be directed at the breathing zone, where it
needs to penetrate the free convection flow as well as the flow of exhalation in order to
achieve high quality of inhaled air.

Interaction of airflows
Brohus (1997) investigated the inhaled air quality in regard to a point contaminant source in
a unidirectional flow field generated in a wind channel. He found that the personal exposure
depends highly on the source location as well as the flow direction relative to the person.
Measurements and smoke tests confirm that the thermal boundary layer that entrains and
transports air from below to the breathing zone in a calm environment is considerably
affected at uniform velocities above 0.05 – 0.10 m/s. Melikov and Zhou (1996) showed that
an invading flow with a mean velocity of only 0.1 m/s and a turbulence intensity of 10% is
able to destroy the free convection flow at the neck of a person. The velocity boundary layer
decreased to approx. 40 mm and the temperature boundary layer to less than 10 mm. The air
temperature near the skin surface was decreased by 4°C, which lead to an increase in the
heat flux by 22%.

17
Chapter 1

Cermak et al. (2002) used a two-dimensional particle image velocimeter (PIV) to identify the
complex flow at the breathing zone of a seated person exposed to airflow from a PV system.
A thermal manikin with a realistic breathing function was used to simulate a human being.
The air terminal was the 2nd generation Movable Panel tested by Kaczmarczyk et al. (2002a,
b, accepted) and described in Section 1.3. The panel was positioned in front and above the
breathing zone at a distance of 0.45 m from the manikin’s nose. The PIV system allowed
instantaneous measurements of the velocity field to be taken corresponding to exhalation,
inhalation and pause. Figure 1.4 presents the mean air velocity contours during exhalation
through the mouth when the personalized airflow rate was 5 L/s. The results showed that
the personalized airflow penetrated the free convection flow even at a very low airflow rate.
As documented in the plot, the flow of exhalation was deflected downward and the free
convection flow at the chin destroyed.

Figure 1.4. Distribution of the mean air velocity (m/s). PV airflow rate 5 L/s. Exhalation through the
mouth. Reprinted from Cermak et al. (2002).

Recently, prior to the present study, Cermak and Melikov (2003) examined the interaction of
PV airflow and the room airflow in a mock-up of a 2-person office. The 2nd generation
Movable Panel was combined with an underfloor air distribution system and tested in
regard to the transmission of exhaled air between occupants. Two breathing thermal
manikins facing each other were used. One manikin was polluting and the other one
exposed. They showed that the PV of the polluting manikin increased the concentration of
exhaled air in the room as compared to the reference case of underfloor ventilation alone. As
a result of mediocre efficiency, the PV was not able to protect the exposed manikin from
inhaling contaminants. The inhaled air concentration increased up to 3.6 times when both
the manikins were using their PV systems at 20 L/s per workplace. The study revealed a
possible drawback of the PV concept, but due to the small number of experimental
conditions it was impossible to draw a general conclusion.

The interaction between the airflow generated by PV, occupants (free convection flow
around the body and respiration) and the airflow pattern outside the workplaces has not
been studied in detail. The design of PV as well as the parameters of the supply airflow may
have an impact on the distribution of contaminants at workplaces and in a room. The use of
PV systems by occupants, i.e. their preferred rates, temperatures and directions of supply
airflow that differ at each workplace, may affect the distribution as well.

18
2
2. Objectives

The objective of the present study is to identify the performance of personalized ventilation
in combination with total-volume ventilation in regard to air quality (including transmission
of contaminants between occupants) and thermal comfort in the occupied zone of rooms.
The personalized ventilation design, total-volume ventilation principle and their control are
of main interest.

An additional objective of the study is to recommend a strategy for application of


personalized ventilation in practice.

19
3
3. Method

The performance of personalized ventilation combined with total-volume ventilation was


examined in full-scale experiments. The experiments were divided into two stages according
to the total-volume ventilation system applied (Table 3.1). In the second stage, three sets of
experiments were performed. The method was identical for all experiments with only minor
changes in between the stages.

Table 3.1. Experimental stages.


Stage Total-volume ventilation Aim
1 Mixing, displacement Overall performance
2-1 Underfloor Overall performance
2-2 Underfloor Impact of the throw of UFAD
2-3 Underfloor Impact of the airflow rate of PV

3.1 Experimental design


A mock-up of a typical office with two identical workplaces was built in a climate chamber
(length x width x height = 4.8 x 5.4 x 2.6 m3). Two breathing thermal manikins were used to
simulate occupants (Section 3.5.1). Each workstation consisted of a desk with an air terminal
device for personalized ventilation, a personal computer and a desk lamp (Section 3.4.1).
The workplaces were arranged behind each other with the manikins facing in the same
direction. The layout aimed at creating the worst possible scenario with PV transmitting
large portions of human-produced contaminants from the front manikin to the back
manikin. Figures 3.1 and 3.2 show respectively the layout of the office and its photo. Air
temperature outside the chamber was controlled at 25°C in order to reduce the heat transfer
through the walls. The office thus simulates an interior section of an open-plan office, if the
blocking effect of the walls is disregarded. The manikins were dressed with underwear,
short-sleeved T-shirt, pants, socks and shoes, giving a total clothing insulation of 0.45 clo
(estimated). The upholstered office chairs on which the manikins were seated had an
additional thermal insulation of 0.15 clo.

Floor covering and occupants with their bioeffluents and exhaled air were selected as the
most common contaminant sources in offices. Three different tracer-gases were used
(Section 3.4.2). The floor covering was simulated by means of a rectangular grid of tubing,
from which carbon dioxide (CO2) was released over the entire floor area. The tubing was
perforated every 0.6 m creating an array of 8x8 dosing points. Air exhaled from the front
manikin was marked with sulphur hexafluoride (SF6). The bioeffluents were simulated with
nitrogen dioxide (N2O). The gas was released at three locations under the clothing of the
front manikin. Both the floor covering and the exhaled air from the front manikin were
present in all the experiments (both stages). The bioeffluents were studied only in stages 1
and 2-1. In the stage 2-2, N2O was used to mark the air exhaled from the back manikin (such
as SF6 for the front manikin). This allowed for the evaluation of the transmission of exhaled

21
Chapter 3

air between the two manikins. In stage 2-3 different contaminant sources were used (see
Table 5.3).

Two types of air terminal device for PV were tested. They were the Round Movable Panel
(RMP) mounted on a movable arm duct and the Vertical Desk Grill (VDG) positioned at the
front edge of a desk (Section 3.3.1). Both types currently represent the most promising
solutions for providing excellent air quality and preferred thermal comfort for occupants.
The major difference is the direction of personalized airflow in respect to a person. The flow
from the RMP and the VDG is respectively transversal and assisting to the flow in the
human boundary layer. Only one type of terminal was tested at a time. The positioning was
adjusted in order to comply with the positioning most often preferred by people
(Kaczmarczyk et al., 2002b; Kaczmarczyk, 2003). The positioning did not change during the
study.

Figure 3.1 Office plan: (1) front thermal manikin – 23 sections, (2) back thermal manikin – 16
sections, (3) Personalized ventilation – RMP, (4) Personalized ventilation – VDG, (5) Displacement
ventilation supply, (6) Mixing ventilation supply, (7) Underfloor ventilation supply, (8) Exhaust, (9)
17” computer monitor – 70 W, (10) Computer tower – 75 W, (11) Desk lamp – 55 W, (12) Ceiling
light fixture – 6 W, (A)-(E) Measurement positions.

Four scenarios were examined: (1) both manikins using PV, (2) front manikin using PV
while back manikin did not or (3) vice versa, (4) neither of the manikins using PV, i.e. total-
volume ventilation alone (reference case). In most cases the personalized airflow rate was
either 7 L/s or 15 L/s per person. In one experiment 30 L/s per person was supplied. The
supply air temperature was 20°C in all experiments (all stages). The combinations tested are
presented in detail at the beginning of each result section.

The performance of PV was studied in combination with mixing, displacement and


underfloor ventilation (Table 3.1). Mixing and displacement ventilation systems are well
understood, besides the fact that they are the most common air distribution principles used
today. The underfloor ventilation system was studied more extensively. An installation of
PV in conjunction with a raised floor is easy, allowing both PV and total-volume systems to
benefit from the same platform. For all combinations air was exhausted at the ceiling level.

22
Chapter 3

Figure 3.2. Photo of the office. The arrangement is identical to that used in the study except: (1) the
manikins did not have wigs and were dressed in short-sleeved T-shirts and (2) there were two racks
holding low velocity anemometers and tracer-gas sampling tubes in the room.

The number and distribution of the heat sources corresponded to a typical office. The total
sensible heat gain was 580 W (22.5 W/m2) and constant during all experiments. Studies
showed (Kaczmarczyk et al., accepted; Zeng, 2002; Yang, 2003) that the improvement of
inhaled air quality and thermal comfort with PV is greater at elevated room air temperature.
The exhaust air temperature was designed at 26°C. The temperature in the occupied zone
was intended to fulfil the requirements of present standards for summer conditions (ISO,
1994; CEN, 1998 – category B). Both PV and total-volume ventilation supplied outdoor air.
The amount of outdoor air per person was 40 L/s in most experiments, and down to 25 L/s
in some experiments in stage 2-2 (see below). The ventilation rate fulfilled the requirements
for removal of sensory pollution load when 40-100% tobacco smokers are present, according
to CR 1752 (1998), category B, and ensured high elevation of interfacial layer (stratification
height) when displacement ventilation was used. The recirculation of exhaust air to the
supply air was not utilized in order to increase sensitivity of the tracer-gas measurements.

The supply air temperature and airflow rates were set in order to allow for comparison of
the combined systems. In stages 1 and 2-1 the supply air temperature was 20°C. Assuming
ideal mixing, the total airflow rate was calculated at 80 L/s (= 4.3 air changes per hour). The
total amount of air supplied to the office was kept constant. The use of PV thus led to a
decrease in the total-volume ventilation rate (i.e. VAV control). In the stage 2-2 experiments
the impact of the throw of UFAD on the performance of the combined systems was
examined. Two throw heights were achieved by providing different airflow rates. The first
airflow rate was 80 L/s, i.e. identical to the previous stages. The second rate was 50 L/s (=
2.7 air changes per hour). In the reference cases without PV, the supply air temperatures of
UFAD were respectively 20 and 16.4°C. Contrary to stages 1 and 2-1, the underfloor airflow
rate was kept constant (i.e. CAV control), 80 L/s or 50 L/s, regardless of the PV use. When
PV was used, the supply air temperature of underfloor ventilation was increased in order to
ensure a constant cooling capacity of ventilation air (PV combined with UFAD). The system
was not under thermostatic control and the supply air temperature set point was
determined from a calculation. A detailed overview of the supply air conditions is given in
Table 5.2, page 74.

The performance of the combined systems was studied in regard to inhaled air quality and
thermal comfort. Section 3.7 describes the criteria for evaluation. The breathing thermal

23
Chapter 3

manikins were used for the assessment of inhaled air quality (contaminant concentration,
inhaled air temperature) and thermal environment in the workplaces. The assessment of the
environment for standing and/or walking occupants was based on the measurement of
contaminant concentration, air temperature and air velocity at five positions throughout the
room (Figure 3.1). At positions A and B the measurement of air velocity and temperature
were performed at several heights. The supply and exhaust air temperatures and the surface
temperatures of walls were monitored as well.

3.2 Air movement room


Figure 3.3 shows a sketch of the climate chamber. The chamber has a modular structure
based on a load-bearing steel construction with a module of 1.2 m. Its maximum size is 7.2 x
4.8 x 5.0 m3 (length x width x height). The walls surrounding the perimeter are made of
insulated chipboard. One of the walls is single glazed up to a height of 2.4 m. The floor is
made of 0.6 x 0.6 m2 chipboard tiles raised 0.7 m above a structural concrete slab. The surface
of the floor is finished with a low-polluting material. The ceiling is made of 0.6 x 0.6 m2
gypsum tiles suspended 0.4 m from the ceiling steel construction. The ceiling is divided into
3 parts (2.4 x 4.8 m2) and can be adjusted to different heights, each part separately or as a
whole. A height of 2.6 m was used in the present study. The chamber size was also modified
with a floor-to-ceiling partition, which splits the chamber into two sections. The larger
section, sized 5.4 x 4.8 m2, was used in the experiments, while the other part was empty and
not used. Neither the ceiling nor the floor was thermally insulated. There were six
fluorescent light fixtures evenly distributed along the ceiling in order to provide overall
lighting. The chamber, including the light fixtures and the door, was carefully sealed prior to
the experiment.

Figure 3.3 Sketch of climate chamber. The section framed bold sized 5.4 x 4.8 x 2.6 m3 (length x width
x height) was used for the experiment.

The climate chamber was situated in a laboratory hall sized 11 x 10 x 8 m3. The floor and the
ceiling of the hall were made of concrete, the walls of bricks. The hall was ventilated and air-
conditioned by mixing ventilation. The temperature in the laboratory hall was controlled at
25°C, i.e. similar to the mean air temperature inside the chamber.

24
Chapter 3

3.3 Ventilation systems

3.3.1 Personalized ventilation

Round movable panel (RMP)


The RMP was an air terminal device with a round front panel with a diameter of 215 mm.
The face (free) area was 0.027 m2. The RMP was designed so as to provide a low-turbulent
flow with a long potential core in order to reduce mixing of the outdoor air and ambient
room air. The RMP was mounted on a movable arm-duct attached to the desktop. The
construction of the movable arm-duct allowed for changing the position of the ATD and the
angle of the supply jet. Figure 3.4 shows the positioning of the RMP used in the present
study. The figure shows also the posture of a manikin and its distances from the desk and
the computer monitor. The layout was identical for the two manikins and maintained
during the whole study.

Figure 3.4 Details of the positioning of the Round Movable Panel

Vertical desk grill (VDG)


The air terminal device consisted of a plenum box with two air discharge openings. The
plenum box was attached underneath the desktop. The supply air entered the plenum at the
back of the desk. The two discharge openings were located in the horizontal and the vertical
plane at the front edge of the desk. The opening on the horizontal plane (20 x 220 mm2)
supplied air vertically to the breathing zone. The opening in the vertical plane (15 x 245
mm2), which was aimed at the horizontal supply toward a torso, was closed. The openings
were equipped with vanes for directing the airflows. Figure 3.5 shows the construction and
positioning of the vanes as used in the experiments.

Figure 3.5. Vertical desk grill. Construction and positioning of the vanes.

25
Chapter 3

Personalized air temperature and airflow rate


The air terminal devices were provided with conditioned outdoor air. The connection was
made with a flexible duct using the plenum beneath the floor. The supply air was first
cooled in the main air-conditioning units and then heated separately for each workstation to
a temperature of 20°C. The reference temperature sensor was mounted direct in the air
terminal. The on/off control used caused fluctuations of the supply air temperature by
approx. ±0.1°C around the mean. The air humidity was not controlled.

A differential pressure sensor (Micatrone, type MFS-C-080) was used to measure the airflow
rate. The airflow rate was proportional to the pressure difference measured before and after
the body of the sensor. The sensor was calibrated against an orifice plate and with a tracer-
gas constant emission method prior to the experiments. The construction and characteristics
of the sensor ensured that it was possible to measure flows down to 3 L/s with high
accuracy and low drop of pressure. The measurement error as specified by the sensor
manufacturer was < 3% of the actual airflow. The pressure differential was measured with a
high precision micromanometer (Furness, type FCO510). The accuracy was 0.025% of
reading up to 20 Pa, ± 0.01 Pa; and 0.25% of reading between 20 and 200 Pa, ± 0.01 Pa. The
measured pressure was roughly 4, 17 and 70 Pa at a rate of respectively 7, 15 and 30 L/s.

3.3.2 Total-volume ventilation


Commercially available supply and exhaust air terminals were used. The terminals were
sized based on the manufacturers’ guidelines. The positioning of the air terminals in the
room is shown in Figure 3.1.

Mixing air terminal device


A swirl type diffuser (TROX, type RFD-R-D-US/250), shown schematically in Figure 3.6,
was used for its ability to ensure a high level of induction and rapid reduction in
temperature differential. The diffuser was suitable for the VAV application. Its performance
was reportedly optimal between 30 and 110 L/s (i.e. 25 to 100% of the max. airflow rate).

Displacement air terminal device


A semicircular unit (Lindab, type COMDIF-CBA 2010) with a radius of planar projection of
250 mm and a height of 1000 mm was used for displacement ventilation (Figure 3.8). The
unit was fitted with nozzles, which made it possible to change the geometry of the near
zone. The adjustment ensured that the supply airflow spread mainly along the walls and
only minimally perpendicular to the room. The near zone, defined as a horizontal distance
from the wall to the place in a room where the maximum velocity decreases to 0.2 m/s, was
predicted and experimentally verified not to be longer than 0.7 m.

Underfloor air terminal device


Four swirl diffusers (TROX, type FBM-3-EU-K/200-SM) were used for underfloor
ventilation (Figure 3.7). The supply ductwork was balanced in order to provide an equal
amount of air through each diffuser. The airflow rate through each diffuser was measured
with a flow capture hood meter. Each diffuser was fitted with a swirl element, which
allowed for adjustments of the swirl in a vertical or a horizontal (radial) direction. The
diffuser core had a diameter of 200 mm. The size of the diffusers and their number were
selected according to a predicted throw height. Section 5.1.3 presents the results of the air
velocity measurement and smoke visualization of the air discharge pattern. The prediction
agreed with the measurement.

26
Figure 3.6 Swirl diffuser for mixing ventilation.
Dimensions in mm.

Figure 3.8. Semicircular air


distribution unit for displacement
ventilation.

Figure 3.7. Floor diffuser installed in a floor panel: (1)


diffuser core, (2) swirl element for adjustment of air
discharge direction, (3) dirt trap, (4) plenum box.

Exhaust air terminal device


The exhaust air terminals were four circular ceiling diffusers (∅D = 160 mm) with a
perforated front plate (Lindab, type PCA-160). The exhaust airflow rate was also balanced
between the diffusers.

Supply and exhaust air temperature and airflow rate


The supply and exhaust air temperatures were measured with a thick film thermistor
mounted direct in the terminal device. The thermistors were not radiation shielded;
however, they were mounted with care in order to prevent the influence of radiation from
the room. Only one sensor was used for the four floor diffusers. The exhaust air temperature
was measured with two sensors on the diagonal (the difference was small). The sensors
were calibrated against a precision mercury thermometer with a scale division of 0.1°C prior
to the experiments. The uncertainty of temperature measurement was estimated to ±0.3°C
with a high level of confidence.

The supply and exhaust airflow rate were determined by means of an orifice plate
measuring section. The installation complied with ISO (1991). Two sizes of an orifice plate
were used to cover the range from 50 to 110 L/s. An industrial transducer (Halstrup-
Walcher, type PU-0.5-S-230-X-L-02) with an accuracy of 0.5% of reading (0-500 Pa) was used
to measure the pressure differential. The accuracy was estimated based on manufacturer’s
data and checked with a reference manometer, which was used for the measurement of a
personalized airflow rate (Section 3.3.1). The accuracy of the airflow rate measurement was
better than 2% (safe estimate).

27
Chapter 3

3.4 Heat and contaminant sources simulation

3.4.1 Heat sources


Table 3.2 shows the total sensible heat gain of the equipment used in the simulated office
mock-up. The electrical power consumption was measured and assumed to be equal to the
total heat gain (radiant plus convective). The accuracy of the consumption meter as reported
by a manufacturer was 0.4 W. The layout of the sources is shown in Figure 3.1.

Table 3.2. Total sensible heat gain of the heat sources.


Source Sensible heat gain (W)
Breathing thermal manikin 2x 75
Computer tower 2x 74
Computer monitor (CRT, 17") 2x 70
Desk lamp 2x 55
Ceiling light fixtures 6x 6
Total 584 (= 22.5 W/m2)

3.4.2 Contaminant sources


Three contaminant sources were simulated by means of a constant emission of different
tracer-gases. The emission rate was determined considering the density (impact of
buoyancy), safe health exposure, measurement range and accuracy of the gas analyzer and
consumption of the gas in respect to its price. Table 3.3 provides a summary of the sources,
characteristics of the gases and the amounts used. The doses as well as the exhaust air
concentration are given in their typical ranges, because the doses differed between the
experimental stages.

The dosing set-up was identical for the three gases. It consisted of a gas cylinder, 2-step
reduction valve, a needle valve (to adjust the resistance of the distribution tubing) and a
glass tube flowmeter (rotameter). The flowmeter was used to monitor the flow stability (not
for measurement). The flow rate (dose) was determined from a mass balance of the tracer-
gas (i.e. supply rate of outdoor air multiplied by the concentration differential between the
exhaust and the supply). The whole arrangement was placed outside the chamber. Practical
experience showed that the dosing was constant during the experiments.

Table 3.3.Summary of the contaminant sources.


Source Tracer-gas Density* Dose** Supply air Exhaust air
(kg/m3) (mL/s) (ppm) (ppm)
Floor covering CO2 1.83 24 - 40 400 700 - 900
Exhaled air (front m.) SF6 6.07 0.14 - 0.18 0.01 1.8 - 2.2
Exhaled air (back m.) N2O 1.83 0.12 - 0.15 0.3 1.8 - 2.2
& Bioeffluents
* at 20°C and 101325 Pa
** at 80 L/s (= 4.3 air change per hour)

28
Chapter 3

Floor covering
In order to achieve a uniform distribution over the entire floor a 4-quadrant symmetrical
grid of PVC tubing was arranged on the floor of the chamber (Figure 3.9). Very small holes
were pierced into the tubing forming a grid of dosing points with distance between the
points of 0.6 m. The tubing was placed in metal U-profile beds mounted in between 16 mm
thick chipboard plates. A distribution box was arranged in the intersection of the quadrants
(i.e. in the centre of the room) and connected with the flowmeter outside the chamber. The
CO2 temperature in the distribution box was monitored in order to make sure that the gas
was discharged at the room temperature. The discharge temperature was not controlled.
Any impact of the expansion of the tracer-gas (after pressure reduction) on the gas
temperature was not observed.

Figure 3.9. Simulation of floor contaminant source. Left: one quarter of the floor area with a
distribution box in the centre of the room. Right: details of the chipboard and positioning of the tubing
in a U-profile spacer. Dimensions in mm.

High resistance of the tubing perforation (and therefore a high overpressure in the system),
in contrast to a low flow resistance of the tubing, is believed to have ensured the uniformity
of the gas distribution. The testing of the uniformity could be troublesome. In the present
study it was examined based on the concentrations measured at numerous locations
throughout the chamber. The chamber was ventilated with 50 L/s using either the mixing or
the displacement principle. The distribution was considered uniform because the
concentrations measured at several locations were identical.

Exhaled air
Two different tracer-gases were used to mark the air exhaled from the two manikins. SF6
was used for the front manikin in all experiments, while in stage 2-2 experiments N2O was
switched from simulating bioeffluents generated from the front manikin (see below) to mark
the air exhaled from the back manikin. The dosing set-up was identical. After the flowmeter

29
Chapter 3

there was a 3-way solenoid valve mounted inside each lung. The valve was synchronized
with breathing so that the gas was released to the exhalation airway only when the
exhalation took place. Providing the pulmonary ventilation (breathing rate) was 6 L/min,
the exhaled air contained 1400 – 1800 ppm of SF6 or 1200 – 1500 ppm of N2O (Table 3.3). The
exhaled air was heated to respectively 36 and 34.3°C in order to have a density similar to
that of air exhaled by people (Section 3.5.1). The heating compensated also low water
content in the exhaled air (~ 15% RH exhaled).

Human bioeffluents
The front manikin was also used as a source of human bioeffluents. A constant dose of N2O
was released at three points under the clothing of the manikin: two points were located at
armpits and one at a pelvis region. The dosing set-up was identical to the set-up for the
exhaled air, except for the 3-way valve (inside each lung). The discharge temperature of N2O
was similar to the surface temperature of the manikin, as it was heated when circulated
under the clothing.

3.5 Measuring Instruments

3.5.1 Breathing thermal manikin


Two breathing thermal manikins were used to simulate occupants. The manikins were
identical, except the number of segments their surface was divided into: the front manikin
consisted of 23 sections and the back manikin consisted of 16 sections. The second, less
important difference, was the location of the connectors for supply and exhaust of
respiration air. The connectors were placed at the waist and on the top of the head for the
front manikin and the back manikin, respectively. Table 3.4 lists the manikins’ body
segments and their surface area.

The manikins are shaped as a 1.7 m tall average woman. Their body is made of a 3 mm
fiberglass armed polystyrene shell. The junctions at neck, shoulders, hips and knees allow
the body to be adjusted in a variety of postures. The surface of the manikins is divided into
several sections, each independently heated by means of electric resistance wires. A 0.5 mm
thick cover of a glass fiber shield protects the wiring from mechanical damage. The distance
between the wires is less than 2 mm in order to ensure uniform temperature distribution on
the surface.

Thermal control
The surface temperature of each body segment was controlled to be equal to the skin
temperature of an average person under thermal neutrality. The control is based on the
correlation between the skin temperature and the dry heat loss of an average human body
according to Fanger’s comfort equation (Tanabe et al., 1994):

Tsk = 36.4 − 0.054Qt (3.1)

where Tsk is the skin surface temperature, °C,


Qt is the sensible heat loss, W/m2,
36.4 is the deep body temperature, °C,
0.054 is thermal resistance offset of the skin temperature control system, °C.m2/W.

30
Chapter 3

The control system adjusts the power (power consumption is equal to the sensible heat loss)
in order to fulfil Equation 3.1. The surface temperature for each body segment is determined
from the resistance of the wires. The temperature-resistance relationship was calibrated
prior to the experiments.

Table 3.4. Surface area of manikins’ body segments (m3).


Body segment Front manikin Body segment Back manikin
Left foot 0.043 Left foot 0.043
Right foot 0.043 Right foot 0.041
Left lower leg 0.090 Left lower leg 0.089
Right lower leg 0.090 Right lower leg 0.089
Left front thigh 0.080 Left front thigh 0.16
Left back thigh 0.080 Left back thigh 0.165
Right front thigh 0.083
Right back thigh 0.083
Pelvis 0.055 Pelvis 0.182
Back side 0.110
Scull 0.050 Head 0.1
Left face 0.026
Right face 0.026
Back of neck 0.025
Left hand 0.038 Left hand 0.038
Right hand 0.037 Right hand 0.037
Left forearm 0.050 Left forearm 0.052
Right forearm 0.050 Right forearm 0.052
Left upper arm 0.073 Left upper arm 0.073
Right upper arm 0.078 Right upper arm 0.078
Left chest 0.070 Chest 0.144
Right chest 0.070
Back 0.130 Back 0.133
All 1.480 All 1.476

3.5.2 Artificial lung


Each manikin was equipped with an artificial lung that simulates the human breathing
function (Melikov et al., 2000). The lung is placed outside the manikins and the respiration
air is delivered to and from the manikins with flexible tubing. The breathing cycle
(inhalation, exhalation and pause) and the amount of respiration air as well as the
temperature and humidity of the exhaled air could be controlled. In the present study the
lung was adjusted to simulate breathing of an average sedentary person performing work of
light physical activity.

The breathing cycle consisted of 2.5 s inhalation, 2.5 s exhalation and pause. The pause of the
front manikins and the back manikin was set to last for 0.9 s and 1.1 s, respectively, in order
to prevent their synchronization. The breathing frequency was 10 per minute and the
pulmonary ventilation was 6 L/min, or 0.6 L per breath. The instantaneous rate was higher,
however, because both inhalation and exhalation took 2.5 s of the 6 s breathing cycle, hence:
0.6 L / 2.5 s = 0.24 L/s = 14.4 L/min. The pulmonary ventilation was monitored by means of
a glass tube flowmeter (rotameter). The measurement accuracy was ±1.1 L/min
(corresponds to ±5% of full scale).

31
Chapter 3

The exhaled air was not humidified in the present study, but heated to a density that is close
to the density of air exhaled by people. The density was calculated to be 1.144 kg/m3 based
on the following exhaled air properties: air exhaled from a person consists of 78.1 vol.% N20,
17.3 vol.% O2, 3.6 vol.% CO2 and 0.9 vol.% of Ar, its temperature is approximately 34°C at
room air temperatures between 20 and 26°C (Höppe, 1981), the relative humidity is close to
95%. However, in the experiments, the actual temperature of air exhaled from the two
manikins was different and not identical, due to the different tracer-gas contents (see Section
3.4.2).

The breathing openings are shaped so as to mimic a real person. Providing the manikin sits
upright, the two jets emerging from the nose are declined 45° from the horizontal plane, and
30° from each other. Each of the nostrils has a diameter of 8 mm. The oval shaped mouth
distributes the exhaled air horizontally. The width and the height of the mouth opening are
25 and 5 mm, respectively. Although both the nose and the mouth can be used for
exhalation and inhalation, the air was exhaled through the nose (most typical breathing
pattern) and inhaled through the mouth. The reason for using the different openings is that
there would be a shortcut of exhaled air containing a tracer-gas to the inhaled air (where the
gas concentration is analysed) if the same opening were used.

Figure 3.10. Temperature sensor mounted in the mouth of a manikin.

Inhaled air temperature


The temperature of air inhaled to the manikins was measured with a fast response
Thermobead sensor Series B07, with a time constant of 0.12 s in still air. The sensors were
mounted in the mouth of the manikins (Figure 3.10) and connected by a transducer to a A/D
card (personal computer). Both sensors were calibrated simultaneously against a precision
mercury thermometer with a scale division of 0.1°C prior to the experiments. The
uncertainty of the measurement was estimated at ±0.2°C with a high level of confidence.

3.5.3 Tracer-gas analyzer


The concentration of the three tracer-gases, used to simulate pollution in the present study,
was measured continuously with a multi-gas monitor (Brüel & Kjær, type 1302) based on the
photo-acoustic infrared detection method of measurement. The measurement range was 1.5
- 150000, 0.004 - 400 and 0.03 – 3000 ppm for CO2, SF6 and N2O, respectively. The analyzer
was calibrated for all gases in a certified laboratory prior to the experiments.

A multipoint sampler unit (Brüel & Kjær, type 1303) was used to deliver samples of air from
up to 6 locations at a time to the analyzer. The samples of air were exhausted outside the
laboratory in order to prevent contamination. Both the analyzer and the sampler were
controlled with a personal computer, which was also used to store the measured data.

32
Chapter 3

3.5.4 Low velocity anemometers


Instantaneous values of velocity and
temperature were measured
simultaneously at several locations
throughout the room. The system
(Sensor, type HT400) consists of 16
omni-directional thermal anemometers
(Figure 3.11) connected to a
measurement station. The velocity
sensor was spherical with a diameter of
2 mm, ensuring a fast response. The
temperature sensor was shielded
against radiation. The temperature
compensation of the velocity
measurement was utilized. Table 3.5
presents the technical specification. Figure 3.11. Low velocity thermal anemometer.

Table 3.5. Technical specification of low velocity anemometers.


Measurement velocity range 0.05 to 5 m/s
Repeatability range of 0.05 to 1 m/s ±0.02 m/s, ±1 % of reading
range of 1 to 5 m/s ±3 % of reading
Accuracy of temp. compensation better than ±0.2 %/K
Upper frequency* min. 1 Hz, typ. 1.5 Hz
Temperature range -10 to +50 °C
Accuracy of temp. measurement 0.3 °C
* The upper frequency is defined as the highest frequency up to which the standard
deviation ratio remains in the limits of 0.9 to 1.1 (Melikov et al., 1998)

3.6 Procedure
All experiments were performed under steady-state conditions. Each experiment lasted for 1
day. The procedure as well as data analyses were designed in order to reduce the impact of
possible instability of the process in time.

3.6.1 Concentration measurement


The tracer-gas concentration was measured in the air inhaled by the two manikins, at 5
positions throughout the room as shown in Figure 3.1, in the supply and exhaust of the
chamber, and in the laboratory hall where the test room was located. The concentrations
were measured at heights of 0.1, 0.6, 1.1, 1.4, 1.7 and 2.2 m at positions A and B, and at a
height of 1.7 m at positions C, D and E. The positions were selected in order to characterize
the room airflow pattern. The A and B positions aimed at describing the environment in the
vicinity of the workplaces in terms of concentration gradients. The C, D and E positions
represented the inhaled air quality for a walking person. In total, there were 20 locations
measured.

The supply and exhaust airflow rates and the supply air temperature of both personalized
and total-volume (TV) ventilation were adjusted in the evening prior to an experiment. This
allowed for creating a steady-state condition in the chamber overnight. The dosing of the

33
Chapter 3

tracer-gases started in the morning next day. The tracer-gas distribution was allowed to
reach steady-state conditions in approx. 3 hours. The measurement locations were grouped
and measured in sequences of 6 (Table 3.6 and Table 3.7). The samples in each sequence
were analysed one after another in a loop. The time required to analyse a sequence of 6
channels was about 10 minutes, i.e. the concentration readings from a given location were
available in 10 minute intervals. Each location was sampled at least 10, but usually 12 times
(~ 2 hours) in order to make reliable statistical analyses possible. In each A1 sequence (see
below) 3 readings were typically acquired from each channel (~ 30 minutes). The dosing
system was turned off at the end of the experiment and the conditions were set for another
day. All heat sources in the chamber including the thermal manikins were running all the
time.

Table 3.6. Measurement sequences during stage 1 and 2-1 experiments.


Channel Seq.A1 Seq.A2 Seq.A3 Seq.A4
1 Laboratory Lung manikin 1 Pos. A (0.1 m) Pos. B (0.1 m)
2 Laboratory Lung manikin 2 Pos. A (0.6 m) Pos. B (0.6 m)
3 Supply air (PV) Position C Pos. A (1.1 m) Pos. B (1.1 m)
4 Supply air (PV) Position D Pos. A (1.4 m) Pos. B (1.4 m)
5 Supply air (TV) Position E Pos. A (1.7 m) Pos. B (1.7 m)
6 Exhaust air Exhaust air Pos. A (2.2 m) Pos. B (2.2 m)

Table 3.7. Measurement sequences during stage 2-2 and 2-3 experiments.
Channel Seq.A1 Seq.A2 Seq.A3 Seq.A4 Seq.A5
1 Lung manikin 1 Lung manikin 1 Pos. A - 0.1 m Pos. A - 1.1 m Pos. B - 0.1 m
2 Lung manikin 2 Lung manikin 2 Pos. A - 0.6 m Pos. A - 2.2 m Pos. B - 0.6 m
3 Supply air (PV) Position C Pos. A - 1.1 m Pos. B - 1.1 m Pos. B - 1.1 m
4 Laboratory Position D Pos. A - 1.7 m Pos. B - 2.2 m Pos. B - 1.4 m
5 Supply air (TV) Supply air (TV) Supply air (TV) Supply air (TV) Supply air (TV)
6 Exhaust air Exhaust air Exhaust air Exhaust air Exhaust air

Table 3.8. Order of measurement during stage 1 and 2-1 experiments.


Hours 0 1 2 3 4 5 6 7 8 9 10
Seq.A1 Supply, Exh, Lab
Seq.A2 Inhaled air + C,D,E
Seq.A3 Position A
Seq.A4 Position B
Temperature and velocity

Table 3.9. Order of measurements during stage 2-2 and 2-3 experiments.
Hours 0 1 2 3 4 5 6 7 8 9 10 11
Seq.B1 Supply, Exh, Lab
Seq.B2 Inhaled air + C,D
Seq.B3 Position A
Seq.B4 Position A-B
Seq.B5 Position B
Temperature and velocity

34
Chapter 3

Order of sampling locations


It was desirable to monitor simultaneously the gas concentrations in 8 locations: the supply,
the exhaust and 6 locations at positions A or B. However, the sampling had to be split into
sequences because of the limited number of channels from which the gas analyzer was able
to sample at a time. The order of the sampling locations in sequences (and hence also the
order of the sequences) was different in the two stages. In stages 1 and 2-1, the integrity of
the concentration profiles was prioritized (Table 3.6). The supply and exhaust air
concentrations were measured separately as part of A1 sequence, i.e. in between the other
sequences. At the same time, the laboratory environment was monitored making it possible
to abort the experiment in the case of gas leaking to or from the chamber. The day averages
of the supply and exhaust concentrations were used to calculate the ventilation indexes for
all locations. The order of the sequences is shown in Table 3.8.

The design was improved in stages 2-2 to 2-3. The supply and the exhaust concentration
were monitored continuously during the day (Table 3.7). This was possible because the
measurement of positions A and B was split in 3 sequences.

Table 3.9 presents the order of the sequences. The indexes were calculated using the data
corresponding to the same measurement sequence, thus improving the uncertainty. The
process instability evaluation had also improved because data from a whole day were
available. Detailed analyses performed after the experiments showed that the only source of
instability was the fluctuation of CO2 concentration in the supply air during a day, and that
the uncertainty was practically comparable with the two designs (i.e. orders of sequences
and sampling points).

3.6.2 Temperature and velocity measurements


The room air temperature and velocity, the inhaled air temperature and the heat loss from
the thermal manikins (manikin-based equivalent temperature) were measured several times
during a day at the same occasion. The occasions are indicated in Tables 3.8 and 3.9. Results
showed that the all temperatures were extremely stable during an experiment. Table 3.10
summarizes the sampling frequency and duration of the data acquisition used.

Table 3.10.Sampling frequency and duration of the data acquisition.


Quantity Frequency of Duration of measurement
data acquisition stage 1&2-1/stage 2-2&2-3
Inhaled air temperature 20 Hz 1.5 min
Heat loss from the manikins 3 per minute 3 min/5 min
Room air temperature 5 Hz 3 min/5 min
Room air velocity 5 Hz 3 min/5 min
Boundary conditions 3 per minute 10 min

Inhaled air temperature


The temperature was recorded continuously for 1.5 minutes during the transient breathing.
Only the data corresponding to a period of 2 s during inhalation were selected using a
computer program and analysed. Figure 3.12 shows the typical recording of inhaled air
temperature. The diamonds indicate the samples selected for analyses.

35
Chapter 3

30

Exhalation Pause
29
Temperature (°C)

28

27

26
Inhalation
25
0 5 10 15 20 25 30
Time (s)
Figure 3.12. Inhaled air temperature measurement. The mean inhaled air temperature is 26.1°C.

Heat loss from the manikins


The heat loss from the manikins (Section 3.5.1) was acquired with a personal computer,
which was also used to control the surface temperature of the manikins. Data for all body
segments were stored together with a whole-body average, which was weighted by the
surface area of the segments.

Room air temperature and velocity


The air temperature and air velocity were measured at position A and B using a system of 16
omnidirectional thermal anemometers (Section 3.5.4). The heights of the measurements were
0.05, 0.1, 0.2, 0.6, 1.1, 1.4, 1.7 and 2.2 m in the cases of displacement and underfloor
ventilation and 0.1, 0.35, 0.6, 0.85, 1.1, 1.4, 1.7 and 2.2 m in the case of mixing ventilation.

Boundary conditions
The boundary conditions measurement involved the measurement of the supply and
exhaust air temperatures of PV and total-volume ventilation, laboratory temperature and
surface temperatures of internal chamber walls. The same type of a temperature sensor
(thick film thermistor, described in Section 3.3.2) and a data acquisition unit were used.
There were 4-5 temperature sensors taped on each of the walls, the floor and the ceiling in a
cross-like layout. Results showed that the wall temperature was similar to the air
temperature at a given height, and the results are thus not reported. As regards the supply
air temperature (and total-volume ventilation supply airflow rate), each quantity was
monitored by means of the control system of the air-handling unit continuously during an
experiment. The supply air temperature and rate of personalized air were checked at regular
intervals, corresponding to the other temperature and velocity measurements. Both
quantities were very stable during the experiments.

3.7 Criteria for evaluation


The performance of PV in combination with different total-volume ventilation principles
was examined in regard to the inhaled air quality and thermal comfort. Both seated and
standing/moving occupants were considered in the present study.

The air quality was evaluated based on tracer-gas concentration and air temperature
(Melikov et al., 2000, submitted). Both quantities were measured in the air inhaled by the
breathing thermal manikins and at numerous locations throughout the room. No attempt
was made to derive a combined index, which would e.g. predict the percentage dissatisfied
with the air quality, due to both concentration and temperature. The two quantities were

36
Chapter 3

evaluated separately. Humidity of air, which also affects human perception of air quality,
was not analysed.

The measurement of heat loss from the thermal manikins was used to assess the thermal
comfort of seated occupants. Air temperature and velocity recorded throughout the room
were used for evaluation of thermal comfort outside the workplaces. The criteria for
evaluation included the draught rating and the vertical air temperature difference between
the head and the ankles (ISO, 1994; CEN, 1998).

3.7.1 Contaminant concentration


Normalized concentration
The concentration of contaminants has most often been expressed in term of normalized
concentration, c(–). The concentration was defined as:

c − cS
c(−) = (3.2)
cE − cS

where c is the contaminant concentration in a point;


cS is the contaminant concentration in the supply air;
cE is the contaminant concentration in the exhaust air.

The normalized concentration is equal to 1 if there is complete mixing of air and


contaminants. If the air quality is better than in the exhaust, the normalized concentration is
lower than 1 and vice versa. The supply air has a normalized concentration of 0. The
reciprocal value of the normalized concentration has been referred to as the ventilation
effectiveness, εV (e.g. CEN, 1998).

Personal exposure effectiveness


A personal exposure effectiveness index, εP, was proposed by Melikov et al. (2002) and used
for the evaluation of air terminal devices for personalized ventilation. It was defined as:

c I ,0 − c I
εP = (3.3)
c I , 0 − c S , PV

where cI,0 is the contaminant concentration in the inhaled air without PV;
cI is the contaminant concentration in the inhaled air with PV;
cS,PV is the contaminant concentration in the air supplied from PV.

The personal exposure effectiveness can be interpreted as the portion of clean air from PV in
the inhaled air. It is equal to 1 when 100% of personalized air is inhaled and equal to 0 when
no personalized air is inhaled. The effectiveness of zero thus does not mean that a person
does not inhale any clean air, but only that the person does not inhale any air from PV. The
concept of the index is identical to that of the effectiveness of entrainment in the human
boundary layer introduced by Brohus and Nielsen (1996).

The personal exposure effectiveness can be expressed in terms of the ventilation


effectiveness. The relationship is independent of the contaminant distribution:

37
Chapter 3

cE − cS 1 ε V ,0
εP = 1− = 1− (3.4)
c I ,0 − c S ε V εV

Where εV is the ventilation effectiveness in the inhaled air without PV. In case of mixing (cI,0
= cE) it can be is simplified to:

1
ε P = 1− and ε P = 1 − c ( −) (3.5)
εV

Unlike with ventilation effectiveness, data from two different experiments (with and
without PV) are needed for the calculation of εP in a room where the distribution of a
contaminant is non-uniform. The normalized concentrations were used in order to avoid
error due to a possible difference in the tracer-gas emissions between the experiments. On
the other hand, the uncertainty increased due to the propagation of error.

3.7.2 Temperature
Inhaled air temperature
The inhaled air temperature has been expressed in terms of either absolute temperature (i.e.
as measured) or the normalized temperature TI(–). The normalized temperature was defined
as:

TI − TS , PV
TI ( − ) = (3.6)
TI , 0 − TS , PV

where TI is the inhaled air temperature (with or without PV);


TS,PV is the supply air temperature of PV;
TI,0 is the inhaled air temperature without PV.

The normalized temperature is equal to 1 when PV does not have any impact on the inhaled
air temperature. When the inhaled air temperature equals the supply temperature of PV, the
index becomes 0.

Room temperature distribution


The temperature distribution has been expressed in a similar way, i.e. either in terms of the
absolute temperature or the normalized temperature T(–). The normalized temperature was
defined as:

T − TS
T (−) = (3.7)
TE − TS

where T is the air temperature in a point;


TS is the air temperature in the total-volume ventilation supply;
TE is the air temperature in the room exhaust.

So as for the normalized concentration, the normalized temperature is equal to 1 when the
room air temperature equals the exhaust temperature. The normalized temperature of the
supply air is 0.

38
Chapter 3

Manikin-based equivalent temperature


The manikin-based equivalent temperature, ET, has been used to evaluate the thermal
comfort of seated occupants (Nilsson et al., 1999; Holmér et al., 1999). The manikin-based
equivalent temperature is defined as the temperature of a uniform enclosure in which a
thermal manikin with realistic skin surface temperatures would lose heat at the same rate as
it would in the actual environment (Tanabe et al., 1994). It can be interpreted as a
temperature that a person senses in the actual environment. The relationship between the
equivalent temperature and heat loss for each individual body segment (or the manikin as
whole) can be expressed as:

ET = 36.4 − CQt (3.8)

where ET is the manikin-based equivalent temperature, °C;


C is constant depending on clothing, body posture, chamber characteristics and
thermal resistance offset of the skin surface temperature control system,
°C.m2/W, determined experimentally;
Qt is the sensible heat loss, W/m2.

The constants were determined experimentally. The manikins were exposed to several level
of equivalent temperature in a uniform reference environment (air temperature = mean
radiant temperature, velocity close to zero). The manikins’ clothing and posture were as in
the experiments. The heat loss was measured and the constants calculated.

The constants were then used to determine the equivalent temperature of the actual non-
homogenous environment. Various analyses were carried out:
• The impact of PV on occupants’ thermal comfort was assessed from the change of the
equivalent temperature from the reference condition without PV (∆ET). The segmental
and the whole-body equivalent temperatures were evaluated.
• The vertical temperature difference and the horizontal temperature asymmetry between
(1) front and back of a person and (2) left and right side of a person were used to indicate
the severity of the local thermal discomfort.

3.7.3 Velocity
The velocity distribution was expressed in terms of the mean velocity, v , and the turbulence
intensity, Tu. The turbulence intensity is defined as the standard deviation of the velocity
fluctuations divided by the mean velocity. The velocity and temperature data were used in
the calculation of the draught rating (DR), which was expressed by means of the following
equation (ISO, 1994; CEN, 1998):

DR = (34 − T )(v − 0.05) (0.37 ⋅ v ⋅ Tu + 3.14 )


0.62
(3.9)

3.8 Uncertainty of measurement


Table 3.11 summarizes the typical values of absolute uncertainty based on the analyses of
measurements. The values are given for each uncertainty component together with the
sample uncertainty U and the uncertainty of a derived quantity Uc. The uncertainty
components are described in detail in Appendix A. Because the uncertainty of the
concentration measurements was largely influenced by the component Umeas, a single typical

39
Chapter 3

value cannot be given. The instrument uncertainty Uinstr was the strongest component in the
case of other quantities, and the uncertainty of a measured quantity U can be thus
considered a constant for practical purposes. Except for the concentration measurement,
which varies among the measured locations, the uncertainty of the temperature and velocity
measurement is generally not presented together with the mean value in the result sections.
When presented, the uncertainty is indicated by means of error bars. The level of confidence
is 95%.

Table 3.11. Typical values of absolute uncertainty with a level of confidence of 95%. The component
having the largest impact on the combined uncertainty is printed in bold.

Quantity Umeas Ustab Uinstr U Uc


Fluctuations
Concentration due to nature, 1% of
(inhaled and room) 10-12 readings Stable* reading** See results See results
Inhaled air <0.03°C, TI(-):
temperature 560 readings <0.05°C 0.2°C 0.23°C < 0.07***
Manikin-based < 0.05°C,
equivalent temp. 15 readings < 0.01°C 0.2°C 0.23°C ∆ET: 0.33°C
Surface < 0.03°C,
temperature 30 readings < 0.01°C 0.3°C 0.31°C −
Room air
temperature − < 0.01°C 0.3°C 0.3°C T(-): < 0.1***
Standard
Room air velocity deviation or Tu < 0.005 m/s < 0.025 m/s < 0.025 m/s −
* ca. ±3% fluctuations of supply CO2 due to weather, otherwise very stable
** rectangular distribution
*** uncertainty decreases as the derived index decreases

40
4
4. Personalized, mixing and displacement
ventilation

4.1 Objectives
The objective of the work presented in this chapter is to investigate the performance of PV in
combination with the total-volume ventilation principles most used in practice: mixing and
displacement ventilation.

4.2 Experimental conditions


The two types of ATDs for a personalized ventilation system, namely Round Movable Panel
(RMP) and vertical desk grill (VDG), were combined with an overhead mixing ventilation
system and a displacement ventilation system. A variable air volume strategy was used for
the control of the total-volume ventilation. The contaminant sources examined were a floor
covering, and bioeffluents and exhaled air produced by the front manikin. In the following,
the front manikin and the back manikin are referred to as the polluting manikin and the
exposed manikin, respectively. Several scenarios of PV use in terms of the PV supply airflow
rates were tested. They are summarized in Table 4.1. The supply temperature for both PV
and total-volume ventilation was 20°C and constant in all experiments, aiming at an exhaust
air temperature of 26°C. The experimental design is described in detail in Chapter 3.

Table 4.1. Combinations of personalized airflow rate (L/s) tested.


PV type Mixing ventilation Displacement ventilation
Polluting Exposed Polluting Exposed
manikin manikin manikin manikin
- 0 0 0 0
RMP 0 7 0 7
RMP 0 15 0 15
RMP 15 0 15 0
RMP 15 7 15 7
RMP 15 15 15 15
VDG 0 7 0 7
VDG 0 15 0 15
VDG 15 0 15 0
VDG 15 7 15 7
VDG 15 15 15 15

4.3 Visualization of personalized airflow


A smoke visualization was used to reveal the local airflow pattern in front of the thermal
manikins exposed to personalized air. Figure 4.1 presents a smoke visualization at airflow

41
Chapter 4

rates of 7 L/s and 15 L/s per person. At the rate of 15 L/s, the clean personalized air reached
the face of the manikin directly with both the terminals. At the lower rate the buoyancy
forces affected easily the weak airflow from the RMP terminal – the supplied air dropped on
the desk and mixed with the surrounding polluted air. The face velocity (airflow rate
divided by the free area of the outlet) was 0.26 m/s at the rate of 7 L/s. Reynolds and
Archimedes numbers were 3200 and 0.43, respectively. It is possible that personalized
airflow could accelerate, due to its negative buoyancy, and penetrate the free convection
flow even at a lower rate, if rearranged (Bolashikov et al., 2003). Supplying personalized air
along the free convection flow around the body should make it easier for the VDG to
achieve high air quality in the breathing zone. However, the flow from the VDG was
relatively strong and turbulent even at a lower rate, entraining large quantities of the
surrounding air. The face velocity was as high as 1.6 m/s (rate of 7 L/s). The free convection
flow with a maximum velocity at the breathing zone height of about 0.25 m/s (Section 1.6)
was most probably destroyed and did not affect the performance of the VDG.

Figure 4.1. Smoke visualization. The RMP terminal (top) and the VDG terminal (bottom).
Personalized airflow rate was 7 L/s per person (left) and 15 L/s per person (right).

4.4 Inhaled air concentration


Figure 4.2 presents the normalized concentration of the three simulated contaminants in air
inhaled by the exposed manikin. The total-volume ventilation principle is mixing. The
concentrations are presented at different PV airflow rates of the exposed manikin, while the
polluting manikin did not use its PV. The concentrations of the three contaminants were
averaged because (due to the mixing) the exposure of the manikin to the three contaminants
was comparable. This is documented in Appendix B. The data in Appendix B also show that

42
Chapter 4

the use of PV at the polluting manikin’s workplace did not affect the inhaled air quality of
the exposed manikin.

1.4
Concentration (C-CS/CE-CS)

1.2 Mixing + RMP


Mixing + VDG
1

0.8

0.6

0.4

0.2

0
0 L/s 7 L/s 15 L/s
Personalized ventilation

Figure 4.2. Concentration of contaminants in air inhaled by the exposed manikin versus PV airflow
rate. The concentration of the floor contaminant, bioeffluents and exhaled air are averaged. The error
bars indicate the maximum uncertainty from the three contaminants with a level of confidence of
95%.

The results in Figure 4.2 show that the use of PV decreased the concentration of
contaminants in the air inhaled by the exposed manikin considerably. The largest
improvement was achieved with the RMP at a rate of 15 L/s. When compared to the mixing
ventilation alone, inhaled air concentration decreased (i.e. the inhaled air quality increased)
as many as 6.9 times. This was comparable to the efficiency of the displacement ventilation
alone measured in this study, which is, however, seldom achieved in practice due to
disturbance to the airflow patter (e.g. activity of occupants). The VDG (at a rate of 15 L/s)
did not decrease the inhaled air concentration as much as the RMP; however, the inhaled air
concentration was still 2.7 times lower than in the case of mixing ventilation alone. At a rate
of 7 L/s the efficiency of the RMP decreased as a result of the interaction of flows in front of
the manikin (cold supply air did not penetrate the free convection flow, dropped on the desk
and mixed with the surrounding polluted air). The inhaled air concentration was lower with
the VDG than with the RMP (at 7 L/s), because the impact of buoyancy forces on the
stronger airflow from the VDG was smaller. Even at a rate of 7 L/s, however, the inhaled air
concentrations provided with both types of PV were lower than the inhaled air
concentrations with mixing ventilation alone.

With PV and displacement ventilation, the exposures to the three contaminants were not
comparable. The details of exposures at different PV airflow rates are documented in
Appendix B. Figure 4.3 presents the concentration of the floor contaminant in the air inhaled
by the exposed manikin with PV and displacement ventilation. The airflow rate of the
exposed manikin’s PV was 7 L/s and 15 L/s. The PV of the polluting manikin was not used.
It is shown that the inhaled air concentration with PV and displacement ventilation was
very similar to the inhaled air concentration with PV and mixing ventilation (Figure 4.2).
The reason was the uniformity of the distribution of the floor contaminant, described in
Section 4.7.1. So as with mixing ventilation, the performance of the two terminals was
different due to their efficiency as well as the interaction of flows around the manikin
(discussed previously).

43
Chapter 4

Concentration (C-CS/CE-CS) 1.4

1.2 Displ. + RMP


Displ. + VDG
1

0.8

0.6

0.4

0.2

0
0 L/s 7 L/s 15 L/s
Personalized ventilation

Figure 4.3. Concentration of floor contaminant inhaled by the exposed manikin with PV coupled with
displacement ventilation.

The inhaled air concentration of exhaled air and bioeffluents was similar in all but one case
(i.e. front manikin using VDG at 15 L/s while back manikin does not, see Appendix B). The
similarities made it possible to average the concentrations of the two contaminants and
consider them as a single contaminant produced by a person. Figure 4.4 plots the averaged
concentrations in the air inhaled by the exposed manikin versus the airflow rate of PV. The
cases when the polluting manikin used and did not use its PV are shown separately.

1.4
Concentration (C-CS/CE-CS)

Polluting manikin
1.2

1 RMP 0 L/s
VDG 0 L/s
0.8 RMP 15 L/s
VDG 15 L/s
0.6

0.4

0.2

0
0 L/s 7 L/s 15 L/s
Personalized ventilation (Exposed manikin)

Figure 4.4. Concentration of exhaled air and bioeffluents (averaged) inhaled by the exposed manikin
with PV coupled with displacement ventilation.

The inhaled air quality of the exposed manikin was fairly high with the displacement
ventilation alone (Figure 4.4, PV airflow rate 0 L/s). The reason was the free convection flow
around the human body, which could transport clean air from the lower levels upward to
the breathing zone. The concentration of the human-produced contaminants inhaled was 6
times lower with the displacement ventilation than with the mixing ventilation. The
measurements showed, however, that the use of PV increased mixing of the human-
produced contaminants generated in its vicinity and thus decreased the inhaled air quality
in the occupied zone. The concentration of the contaminants inhaled by the exposed
manikin, when unprotected by PV, increased 3-4 times compared to the displacement
ventilation alone. Comparison of the terminals shows that the exposure of the unprotected
exposed manikin was about 20% lower with the VDG terminal than with the RMP terminal.
The use of PV by the exposed manikin (PV used for protection) decreased the concentration
of the human-produced contaminants inhaled in a similar way to the decrease of the
concentration of the floor contaminant. The comparison of the total-volume ventilation
systems (Figures 4.2 and 4.4) shows that the inhaled air concentrations were lower with PV

44
Chapter 4

coupled with displacement ventilation than with PV coupled with mixing ventilation. The
differences between the two cases could have been due to the differences in adjustments of
the terminals, which were rearranged between the measurements with mixing and
displacement ventilation. The lowest inhaled air concentration was 0.02, which corresponds
to a ventilation effectiveness of 50 (provided with RMP at a rate of 15 L/s). This was the
highest performance of PV identified in the present study (observed also in some
experiments with PV and underfloor ventilation, Chapter 5).

The inhaled air quality of the exposed manikin was very high as long as PV in the vicinity of
the polluting manikin was not used. Figure 4.4 shows that both types of PV had difficulty in
decreasing the concentration inhaled at a rate of 7 L/s further from the reference case of
displacement ventilation alone. The inhaled air concentration decreased substantially only
with the RMP at a rate of 15 L/s (6 times compared to the displacement ventilation alone).
The use of VDG did not decrease the inhaled air concentration compared to the reference
case due to its low efficiency.

4.5 Inhaled air temperature


The inhaled air temperature was measured in the mouth cavity of each manikin. The
analyses of the results (presented in Appendix C in detail) showed that the operation of the
PV system of one manikin did not change the inhaled air temperature of the other manikin
by more than 0.3°C, regardless of the total-volume ventilation.

28
27 Mixing + RMP
Mixing + VDG
Temperature (°C)

26
Displ. + RMP
25 Displ. + VDG
24
23
22
21
20
0 L/s 7 L/s 15 L/s
Personalized ventilation (Exposed manikin)

Figure 4.5. Inhaled air temperature of the exposed manikin as a function of the personalized airflow
rate. The pollution manikin did not use its PV.

Figure 4.5 presents the inhaled air temperature of the exposed manikin with the mixing
ventilation and the displacement ventilation, alone and in combination with PV. The PV unit
of the polluting manikin was not used. The inhaled air temperature was 26.6°C with the
mixing ventilation alone, and 25.6°C with the displacement alone. The use of PV decreased
the inhaled air temperature from the reference. The lowest inhaled temperature was 20.7°C
with the RMP at a rate of 15 L/s, when combined with the displacement ventilation. The
combination of PV with displacement ventilation consistently provided lower temperatures
of the inhaled air that the combination with mixing ventilation. This is due to the generally
lower temperatures in the occupied zone when the displacement ventilation was applied. It
is more interesting to realize than the relationship between the inhaled air temperature and
personalized airflow rate is similar to the relationship between the inhaled air concentration
and the personalized airflow rate (compare with Figures 4.2 and 4.3). A better air quality,

45
Chapter 4

associated with a lower temperature, was achieved with the RMP at the high airflow rate,
while the VDG terminal performed better at the lower rate.

1.2
Temperature (TI-TS/TI,max-TS)

Mixing
1
Displacement
0.8 y = 1.07x
R2 = 0.96
0.6

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1 1.2
Concentration (CI-CS/CE-CS)

Figure 4.6. Inhaled air temperature versus concentration of a floor contaminant with personalized,
mixing and displacement ventilation. The maximum inhaled air temperature (TI,max) was different for
mixing and displacement ventilation.

Figure 4.6 plots the normalized inhaled air temperature of the exposed manikin versus the
normalized concentration of the floor contaminant with mixing ventilation and
displacement ventilation. It is shown that the two properties are very well correlated. The
relationship presented is independent of the total-volume ventilation principle, because the
floor contaminant concentration was almost uniform with both the displacement and mixing
ventilation (discussed below). The differences in the inhaled air temperature between the
two ventilation principles (temperature was lower with the displacement ventilation than
with the mixing ventilation, Figure 4.5) were eliminated when the maximum inhaled air
temperatures (TI,max) measured for each ventilation principle were used to determine the
normalized air temperature. The correlation between the inhaled air temperature and the
three contaminants with only personalized and mixing ventilation was identical (y = 1.07x,
R2 = 0.96).

1.2
Temperature (TI-TS/TI,max-TS)

0.8

0.6 Mixing
Displacement
0.4 Linear (Mixing)
y = 1.12x
0.2
R2 = 0.96
0
0 0.2 0.4 0.6 0.8 1 1.2
Concentration (CI-CS/CE-CS)

Figure 4.7. Inhaled air temperature versus concentration of exhaled air with personalized, mixing and
displacement ventilation. The maximum inhaled air temperature (TI,max) was different for mixing and
displacement ventilation.

Figure 4.7 presents the relationship between the inhaled air temperature and the
concentration of the exhaled air contaminant with mixing ventilation and displacement
ventilation. The experimental conditions are identical with Figure 4.6. It is shown that the
two properties are well correlated in the case of mixing ventilation, while with displacement

46
Chapter 4

ventilation they are not. The reason was the non-uniformity of the exhaled air concentration
in the room, which was also influenced by the use of PV. The relationship between the
inhaled air temperature and the concentration of bioeffluents was comparable (y = 1.12x, R2
= 0.95).

The correlations presented are not related to the type of contaminant source so much as to
the distribution of the contaminant in a room. A good correlation is obtained when the
distribution is uniform and vice-versa. In rooms with displacement ventilation most
contaminants are distributed non-uniformly. Also the concentrations of the floor
contaminant were literally non-uniform near the floor (Figure 4.16); however, uniformity
was achieved further from the floor in the occupied zone. Therefore, in rooms with
displacement ventilation, the distribution of contaminants must be considered. In rooms
with mixing ventilation, on the other hand, the type and location of the contaminant source
is unimportant.

4.6 Thermal comfort of seated occupants


Equivalent temperature (ET) measured by means of two thermal manikins was used to
evaluate the thermal comfort of seated occupants. As expected, an increase in the PV airflow
rate increased cooling and hence decreased the ET for the exposed body segments and
consequently the ET for the whole body. Figure 4.8 compares the whole-body cooling
performance of PV combined with the mixing ventilation and the displacement ventilation
for the two manikins. PV was used at 15 L/s (both manikins) and 7 L/s (exposed manikin
only). Despite the differences between the ET measured with the two manikins (see below),
the combinations of PV with displacement ventilation are consistently shown to provide a
lower ET that the combination of PV and mixing ventilation. The reason was a lower air
temperature in the occupied zone in the former case (documented in Figure 4.19). On
average, the use of PV at a rate of 15 L/s decreased the ET by 1-2°C. At a rate 7 L/s, the
whole body ET decreased by as little as 0.5°C. The details of the whole-body ET are
presented in Appendix D. The data in Appendix D document that the use of PV by one
manikin did not affect the ET of the other manikin, whether or not its PV unit was used.

27 27
15 L/s 7 L/s Mixing
26 26 Displ.

25 25
ET (°C)

ET (°C)

24 24

23 23

22 22
RMP VDG no PV RMP VDG no PV RMP VDG no PV
Polluting manikin Exposed manikin Exposed manikin

Figure 4.8. Whole-body manikin-based equivalent temperature. Left: both PV units used at 15 L/s per
workplace. Right: polluting manikin using PV at 15 L/s, and exposed manikin using PV at 7 L/s.

It is not certain from the present data which air terminal for PV is able to decrease the
whole-body equivalent temperature the most. The results for the polluting manikin showed
that the RMP terminal provided higher cooling than the VDG, while with the exposed
manikin the classification was opposite. The most likely reason is that the environment

47
Chapter 4

evaluated was not identical, especially with PV. It is assumed that even a slight
misalignment of the PV airflow and/or the difference in the thermal resistance of the
clothing caused by air pockets and folds could have affected the heat loss greatly.

The impact of PV on the segmental ET was analysed in order to reveal possible sources of
local thermal discomfort. Figure 4.9 presents the distribution of ET over the body of the
exposed manikin in the reference cases without PV. It is shown that the distribution of the
ET was different for the mixing ventilation and for the displacement ventilation. The
difference was pronounced most for feet and lower legs (1°C) and less for thighs and pelvis
(0.5°C). The total-volume ventilation principles did not affect the rest of the body.

28

Mixing
27 Displacement
ET (°C)

26

25

24

23
L.Low.Leg
R.Low.Leg

Whole Body
R.Forearm
L.Upper Arm
R.Upper Arm
R.Thigh

Head
L.Hand
R.Hand
L.Forearm
L.Thigh
L.Foot
R.Foot

Pelvis

Back
Chest

Figure 4.9. Distribution of manikin-based equivalent temperature on body segments with mixing and
displacement ventilation alone (without PV).

The cooling effect of PV was expressed in terms of the decrease of the ET measured with PV
from the ET measured without PV in a corresponding reference case. Figures 4.10 and 4.11
present the cooling effect of respectively RMP and VDG on the manikin’s body segments.
The exposed manikin (reported) was using PV at 7 L/s and 15 L/s while the polluting
manikin was using PV at 15 L/s. It is shown that the impact of PV on the body cooling was
very localized (only a few body parts were affected) and strong in comparison with the
impact of total-volume ventilation. The cooling effect of PV was independent of the
background environment, i.e. the decrease of ET from the reference conditions was
comparable for PV in combination with mixing ventilation and for PV in combination with
displacement ventilation. Furthermore, the fact that the ET for most body parts did not
change to any great extent when PV was used indicates that an increase in the ambient air
temperature (recorded near the two workplaces, Figure 4.19) did not have a large impact on
thermal comfort.

48
Chapter 4

-2
∆ ET (K)

-4
Mix. + RMP, 7 L/s
-6 Mix. + RMP, 15 L/s
Displ. + RMP, 7 L/s
-8
Displ. + RMP, 15 L/s
-10
L.Low.Leg
R.Low.Leg

Whole Body
L.Thigh
R.Thigh

Head
L.Hand
R.Hand
L.Foot
R.Foot

Pelvis

L.Forearm
R.Forearm
L.Upper Arm
R.Upper Arm

Back
Chest
Figure 4.10. Cooling effect on manikin’s body segments caused by RMP terminal combined with
mixing and displacement ventilation.

-2
∆ ET (K)

-4
Mix. + VDG, 7 L/s
-6 Mix. + VDG, 15 L/s
Displ. + VDG, 7 L/s
-8
Displ. + VDG, 15 L/s
-10
L.Low.Leg
R.Low.Leg

Whole Body
L.Thigh
R.Thigh

Head
L.Hand
R.Hand
L.Foot
R.Foot

Pelvis

L.Forearm
R.Forearm
L.Upper Arm
R.Upper Arm

Back
Chest

Figure 4.11. Cooling effect on manikin’s body segments caused by VDG terminal combined with
mixing and displacement ventilation.

The parts most affected by PV were the head and chest of the manikin. The greatest cooling
effect was achieved with the VDG, which decreased the ET of the head to as low as 18°C.
This result means that the system may be able to cover large individual differences between
occupants. The gentle cooling of the manikin’s thighs and upper arms indicates that the
cooling of the RMP was distributed more uniformly over the body surface. The distribution
of the ET was similar for the front manikin, which was exposed to only one condition of
personalized airflow rate. The cooling was more uniform at a lower rate of 7 L/s with both
the terminals. With the VDG, however, the cooling still affected hands and chest of the
manikin. It is precisely the cooling of the lower chest that was identified as the major source
of complaint by occupants with VDG (Kaczmarczyk, 2003; Kaczmarczyk et al., 2004).

49
Chapter 4

4.7 Contaminant distribution

4.7.1 Contaminant concentration profiles


In the experiments with mixing ventilation, the concentrations of the floor contaminant,
exhaled air and bioeffluents were uniform along the room height and equal to the exhaust
air concentration. The use of RMP or VDG did not have an impact on the contaminant
distribution. Although uniformity was expected also in the horizontal direction, differences
in magnitude were observed along the length of the room for the exhaled air and
bioeffluents consistently in most experiments. The higher concentrations were measured in
the polluting manikin’s part of the room, i.e. closer to the source, and vice versa. The
normalized concentrations ranged from 1.0 to 1.3 closer to the source (position A), and from
0.8 to 1.0 further form the source (position B). The concentration of the floor contaminant
was horizontally uniform.

With displacement ventilation the contaminant concentrations were stratified. The


distribution of the contaminants was influenced by the personalized air terminal used, its
airflow rate, and the type and location of the contaminant source.

Personalized airflow in the vicinity of the source


The distribution of the exhaled air and bioeffluents changed substantially when the PV of
the polluting manikin was used. Figures 4.12 and 4.13 show the concentration profiles of
bioeffluents and exhaled air measured near the polluting and the exposed manikin,
respectively. The impact of the PV airflow rate on the distribution is compared.

2.5
Displacement alone
Height above floor (m)

2 Displ. + RMP (1)


Displ. + RMP (2)
1.5 Displ. + VDG (1)
Displ. + VDG (2)
1

0.5

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Bioeffluents (C-CS/CE-CS)

2.5
Displacement alone
Height above floor (m)

2 Displ. + RMP (1)


Displ. + RMP (2)
1.5 Displ. + VDG (1)
Displ. + VDG (2)
1

0.5

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Exhaled air (C-CS/CE-CS)
Figure 4.12. Concentration profiles of human bioeffluents (top) and exhaled air (bottom) near the
polluting manikin. Legend: (1) polluting manikin using PV at 15 L/s while exposed manikin does not;
(2) both manikins using PV at 15 L/s.

50
Chapter 4

Height above floor (m) 2.5 2.5 Displacement alone

Height above floor (m)


Displ. + RMP (1)
2 2 Displ. + RMP (2)
Displ. + VDG (1)
1.5 1.5 Displ. + VDG (2)

1 1

0.5 0.5

0 0
0 0.5 1 1.5 2 0 0.5 1 1.5 2
Bioeffluents (C-CS/CE-CS) Exhaled air (C-CS/CE-CS)

Figure 4.13. Concentration profiles of human bioeffluents (left) and exhaled air (right) near the
exposed manikin. Legend: (1) polluting manikin using PV at 15 L/s while exposed manikin does not;
(2) both manikins using PV at 15 L/s.

The contaminant concentration in the occupied zone increased substantially. The profiles
measured in front of the polluting manikin (Figure 4.12) revealed concentration peaks at a
height of 0.6 m. With the RMP the peaks were observed for both the bioeffluents and the
exhaled air. With the VDG a peak was found only for the bioeffluents. The distribution of
the exhaled air had a linear profile without a peak. The peaks can be explained considering
the local airflow pattern, especially the direction of personalized airflow in respect to the
human body. The downward airflow from the RMP transported the contaminants from both
the breathing zone and around the body under the table, in front of which the peaks were
measured (see the smoke visualization presented in Figure 4.1). The airflow from the VDG
allowed the bioeffluents from the pelvis region to spread under the table so that they could
stratify in the lower occupied zone. Therefore, the peak of the bioeffluents with the VDG is
much lower than the peaks with the RMP.

The concentration of the human-produced contaminants increased also next to the exposed
manikin, although there were no peaks detected there (Figure 4.13). The distribution of the
bioeffluents and exhaled air were comparable. In the upper occupied zone the
concentrations increased more with the VDG than with the RMP. This may lead to the
conclusion that VDG causes higher transmission of contaminants between workplaces. The
inhaled air concentration presented in Figure 4.4 showed, however, that the exposure of the
exposed manikin was larger with RMP. The reason must have been a low concentration of
contaminants near the floor, which was transported to the breathing zone by means of the
free convection flow around the body. The differences in the distribution due to RMP and
VDG are less clear from the concentration profiles presented. The comparison of Figures 4.12
and 4.13 illustrates huge spatial differences between the contaminant concentrations. This
suggests that the personal exposures may vary considerably depending on the position of
the person in a room. Moreover, as discussed previously, the inhaled concentration will not
be correlated with the inhaled air temperature.

Personalized airflow not directed towards the source


The ability of PV to affect the distribution of an active contaminant generated from a
localized source located far from a workplace was studied. This was the case of the exhaled
air and bioeffluents released from the polluting manikin. The case is represented with the
exposed manikin using PV while the polluting manikin’s PV was switched off. The analyses
showed (not presented) that unless the polluting manikin used its PV, the distributions of

51
Chapter 4

the bioeffluents and the exhaled air were very similar. The two contaminants were thus
averaged and are presented together as a single contaminant.

2.5 2.5 Close to


exposed manikin
Height above floor (m)

Height above floor (m)


2 2
Displacement alone
1.5 1.5 Displ. + RMP
Displ. + VDG
1 1

0.5 0.5
PV: 7 L/s PV: 15 L/s
0 0
0 0.5 1 1.5 2 0 0.5 1 1.5 2
Concentration (C-CS/CE-CS) Concentration (C-CS/CE-CS)

Figure 4.14. Distribution of human-produced contaminants close to the exposed manikin (position B)
using PV at 7 L/s (left) and 15 L/s (right). PV of the polluting manikin not used.

Figure 4.14 presents the distribution of the exhaled air and bioeffluents with PV and
displacement ventilation measured near the exposed manikin (position B) at two PV airflow
rates. At a rate of 7 L/s, the weak airflow from the PV did not affect substantially the
contaminant distribution in comparison with the displacement ventilation alone, regardless
of the air terminal device used. At a rate of 15 L/s, both the RMP and the VDG caused
mixing in the occupied zone. The profile with the VDG remained linear, but it tilted towards
a higher concentration near the ceiling. The shape of the profile with the RMP may suggest
that there was a two-zonal distribution (characteristic of displacement ventilation). It is more
likely, however, that the profile was influenced by the flow conditions at position B,
considering that the RMP provided clean air to the vicinity of the manikin.

2.5 Close to
polluting manikin
Height above floor (m)

2
Displacement alone
1.5 Displ. + PV: 7 L/s
Displ. + PV: 15 L/s
1

0.5

0
0 0.5 1 1.5 2
Concentration (C-CS/CE-CS)

Figure 4.15. Distribution of human-produced contaminants close to the polluting manikin (position
A), when its PV was not used. Exposed manikin using PV at 7 L/s and 15 L/s.

The analyses showed (not presented) that the type of PV used by the exposed manikin did
not affect the distribution of the human bioeffluents and exhaled air near the other
workplace, where the contaminants were generated. Instead, as documented in Figure 4.15,
the profiles were influenced by the supply airflow rate. Two distinct vertical zones typical
for a displacement ventilation pattern could be identified: a cleaner zone in the lower levels
and a contaminated zone below the ceiling. It may be suggested, however, that the profiles

52
Chapter 4

reflect the presence of the thermal plumes above the polluting manikin’s workplace rather
then the overall airflow pattern. Figure 4.15 shows that the thermal plume carrying
contaminants dissolved higher above the floor when the PV was used, most probably due to
the decrease in the vertical temperature gradient in the room (Figure 4.19).

Figure 4.16 presents the average concentration profiles of a floor contaminant based on the
profiles measured near the two workplaces (positions A and B). The distribution is
compared at different patterns of use of PV. In either case, the personalized airflow did not
encroach directly onto the surface of the floor, i.e. the contaminant source was not affected.

2.5 Displacement alone


Height above floor (m)

Displ. + RMP (1)


2 Displ. + RMP (2)
Displ. + VDG (1)
1.5 Displ. + VDG (2)

0.5

0
0 0.5 1 1.5 2
Floor contam. (C-CS/CE-CS)

Figure 4.16. Average concentration profiles of a floor contaminant based on the profiles measured
near the two manikins. Legend: (1) polluting manikin using PV at 15 L/s while exposed manikin does
not; (2) both manikins using PV at 15 L/s.

The type of PV did not have a substantial impact on the distribution. The distribution was
uniform (~1) above a height of 0.6 m. Despite the contaminant release at the floor level, the
concentration was low near the floor (as compared to the exhaust concentration) because of
the clean air current from displacement ventilation. The free convection flow along a human
body thus transported the clean air upward to the breathing zone. However, substantial
improvement of the inhaled air quality did not materialize due to the entrainment of
contaminated room air from the free convection flow. The results presented in Figure 4.3
showed that the normalized concentration in the inhaled air to the exposed manikin did not
differ by more than 5-10% from unity.

4.7.2 Inhaled air quality of walking occupants


The concentration of a contaminant inhaled by a walking occupant is the same as the
concentration of the contaminant in the ambient air at the breathing zone height at a given
position in the room (Brohus and Nielsen, 1996; Mattsson, 1999). Figures 4.17 and 4.18
present the concentrations of the exhaled air contaminant measured at five positions in the
room at a height of 1.7 m in the case of PV combined with displacement ventilation. The
figures document two scenarios of PV use: the polluting manikin alone using PV at 15 L/s,
and both manikins using PV at 15 L/s. The measurement positions are sorted so as to
represent the exposure of an occupant walking around the desks (Figure 3.1).

53
Chapter 4

Concentration (C-CS/CE-CS) 1.2

0.8

0.6

0.4 Displ. + RMP


0.2 Displ. + VDG
Displacement alone
0
Position: A C D E B

Figure 4.17. Concentration of exhaled air at 5 locations across the room at the breathing zone height
(1.7 m). Only polluting manikin is using PV at 15 L/s.

1.2
Concentration (C-CS/CE-CS)

0.8

0.6

0.4 Displ. + RMP


0.2 Displ. + VDG
Displacement alone
0
Position: A C D E B

Figure 4.18. Concentration of exhaled air at 5 locations across the room at the breathing zone height
(1.7 m). Both PV units are used at 15 L/s per workplace.

With the displacement ventilation alone, the concentrations were generally high in front of
the polluting manikin (position A). Behind the polluting manikin (positions D, E and B) the
concentrations decreased and remained low. The use PV decreased the concentrations in
front of the polluting manikin. However, behind the polluting manikin, the concentration
increased, compared to the displacement ventilation alone. There was a difference between
the two terminals: although the VDG terminal seemed to evacuate the exhaled air
contaminants from the occupied zone, it promoted mixing and increased the concentration
at the breathing zone height more than the RMP. This implies that the use of VDG increases
the exposure of walking occupants to active contaminants. The difference between the
terminals is, however, not large. Figure 4.18 documents that the use of RMP at the exposed
manikin’s workplace increased the concentration in its vicinity to a mixing ventilation level,
and the two terminals became comparable. Moreover, continuous activity of occupants
would destroy the stratification of contaminants and create mixing conditions, as
demonstrated by Mattsson (1999).

4.8 Temperature distribution


The results on the thermal environment measurement showed that the PV did not contribute
substantially to the temperature (and velocity) distribution in the occupied zone. Unlike the
concentration, the temperature was fairly uniform horizontally. Figure 4.19 presents the
vertical air temperature profiles with the mixing ventilation and the displacement

54
Chapter 4

ventilation, alone and in combination with PV used at both workplaces at 15 L/s. The results
of the tests with lower personalized airflow rates, or only a single PV unit, lay between the
profiles of the two cases. The profiles are based on the measurements near the two
workplaces. The repeatability of the temperature measurements was very high.

2.5
Displ. alone
Height above floor (m)

2 Displ. + RMP
Displ. + VDG
1.5 Mixing alone
Mixing + RMP
1 Mixing + VDG

0.5

0
0.2 0.4 0.6 0.8 1 1.2
Temperature (T-TS/TE-TS)

Figure 4.19. Average air temperature profiles based on the profiles measured near the two manikins.
Personalized ventilation airflow rate 15 L/s per workplace.

The use of PV increased the normalized temperature by less than 0.1 near the floor. No
differences between the two terminal types were observed. This suggests that the reason
was rather the decrease in the displacement ventilation airflow rate to 50 L/s than the direct
impact of the PV. The temperatures differentiated for the two terminals between the heights
of 0.6 and 1.4 m. The VDG terminal increased the temperatures more than the RMP. The
largest increase in the dimensionless temperature of about 0.2 was equivalent to an absolute
temperature of 1°C. The differences almost vanished above the occupied zone. With the
VDG and displacement ventilation, the small decrease of the temperature was caused by the
upward direction of the VDG airflow, which affected the upper sensors near the exposed
manikin. The PV combined with the mixing ventilation did not affect the temperature
distribution at any location measured. The distribution was fairly uniform both horizontally
and vertically. The presence of the supply air diffuser on the ceiling caused the decrease of
the temperature above the occupied zone.

4.9 Velocity distribution


The velocity profiles measured near the two manikins were averaged in order to generalize
the results, despite the fact that the two profiles were not absolutely identical. The
differences between the positions were due to the velocity field and not to poor repeatability
or high uncertainty. Figure 4.20 presents the mean air velocity profiles with and without PV.
When applied, both the PV units were used at 15 L/s. The cases with both displacement
ventilation and mixing ventilation are documented in the figure.

With the displacement ventilation, the highest velocity was identified in the vicinity of the
floor due to the horizontal flow emerging from the diffuser mounted on the wall. The
velocities decreased gradually from the floor up to a height of 1 m, above which they
remained low and constant. Only with the VDG, which supplied air upward, did the
velocities increase near the ceiling. The velocities near the floor decreased a little when the
PV was used. This is most probably caused by the decrease in the displacement ventilation
airflow rate from 80 L/s to 50 L/s. However, the uncertainty of measurement should be

55
Chapter 4

considered. The measurement range of the thermal anemometers used starts from 0.05 m/s
with a typical level of uncertainty of the mean of 0.025 m/s (Section 3.8).

2.5
Displ. alone
Height above floor (m)

2 Displ. + RMP
Displ. + VDG
1.5 Mixing alone
Mixing + RMP
1 Mixing + VDG

0.5

0
0 0.05 0.1 0.15 0.2
Velocity (m/s)

Figure 4.20. Average profiles of mean air velocity based on the profiles measured near the two
manikins. Personalized ventilation airflow rate 15 L/s per workplace.

The velocity levels were higher with the mixing ventilation than with the displacement
ventilation. The highest velocities were measured near the floor. The velocity decreased
gradually from the floor, but it increased again near the ceiling. The profile reflects the
airflow pattern. The airflow from the ceiling-based diffuser was deflected from the walls,
directed downward and after being deflected for the second time from the floor it emerged
to the lower occupied zone. The velocities did not decrease below a level of 0.1 m/s in the
reference case. When the PV was used, the velocities decreased down to 0.05 m/s, but most
probably again due to the total-volume ventilation rather than the PV. This is supported by
the fact that the velocities with the personalized and mixing ventilation were always higher
than with the personalized and displacement ventilation, although the PV obviously
contributed to the mixing of the air in both cases.

4.10 Discussion
The two most promising but very different air terminals for PV were combined with the
total-volume ventilation principles most used in practice: mixing and displacement
ventilation. The performance of the combined systems was examined under the most
extreme occupants’ patterns of use. The tested combinations of PV airflow rates reflected the
fact that some occupants may choose not to use the PV system, while others would exploit
the system at its maximum capacity.

The results for the mixing ventilation showed that the distribution of the three contaminants
was very homogenous across the room, although the types and locations of the contaminant
sources were very different. Because of mixing, the concentrations of the contaminants in the
air inhaled by the exposed manikin were comparable and similar to the exhaust air
concentrations (Melikov et al., 2003). As expected, the use of PV improved the inhaled air
quality (i.e. decreased the inhaled air concentration) of the exposed manikin considerably.
As compared to the mixing ventilation alone, the inhaled air concentration decreased 6.8
and 2.6 times in the case of RMP and VDG, respectively. The improvement was similar in
regard to the three contaminants. Hence, PV will always be able to protect occupants from
pollution and thus increase the quality of inhaled air in rooms with mixing air distribution.
The mixing air distribution principle implies that the inhaled air quality does not depend on

56
Chapter 4

the type of contaminant source and its location. The performance of PV tested in the present
study was comparable to the performance of similar air terminal devices reported in the
literature (Bolashikov et al., 2003; Melikov et al., 2002; Faulkner et al., 1999). The inhaled air
concentration and temperature were measured at two rates of personalized air (7 and 15
L/s). Bolashikov et al. (2003) and Melikov et al. (2002) presented the performance
characteristics of respectively RMP and VDG (similar terminal) over a range of airflow rates.

The direction of personalized airflow in respect to the occupant’s head may affect the
inhaled air quality to a great extent. The air terminals tested in the present study complied
with the positioning most often preferred by people in previous experiments (Kaczmarczyk
et al., 2002b) and were not rearranged at different airflow rates. The smoke visualization
demonstrated that cold personalized air may not reach directly the face of an occupant and
thus may not always ensure excellent air quality. In addition, the inhaled air quality might
either increase or decrease if the occupant moves. Hence, unless PV airflow reaches the face
of an occupant directly, the concentration of contaminants in the surrounding ambient air
will affect the inhaled air quality and thus will have to be considered.

In rooms with displacement ventilation the distribution of contaminants is typically non-


uniform, depending on the type and location of the contaminant source. A passive plane
contaminant located on the floor, representing pollution from carpet, PVC or linoleum, was
exposed direct to the clean air current from the wall-based displacement unit. The clean air
was polluted at a very short height (ca. <0.2 m) and the concentration of the floor
contaminant became uniform further from the floor. Although the free convection flow
along a human body transported the clean air upward to the breathing zone, substantial
improvement of the inhaled air quality did not materialize due to the entrainment of
contaminated room air to the free convection flow. The results showed that the normalized
concentration in the air inhaled by the exposed manikin did not differ from unity by more
than 5-10%. The inhaled air concentration was thus not higher than the concentration in the
ambient air, as suggested by Murakami et al. (1998) based on a CFD simulation. The use of
PV did not affect the distribution of the floor contaminant in the room, while the inhaled air
quality of the manikin protected with PV improved several times (Figure 4.3). The
improvement with PV in combination with displacement ventilation (in regard to the floor
contaminant) was comparable to the improvement with PV in combination with mixing
ventilation, i.e. it was independent of the room air distribution principle.

Displacement air distribution is typically associated with a low exposure of occupants to


active contaminants. Also in the present study, the inhaled air quality provided in regard to
the human-produced contaminants was relatively high, equivalent to a ventilation
effectiveness of 6 (Melikov et al., 2003). The measurements showed that it was difficult for
PV providing low airflow rates (i.e. mediocre performance) to improve the already high air
quality. Such high values of air quality may not be typical in rooms with displacement
ventilation in practice. The activity of occupants and flow obstacles may disturb the
buoyancy driven airflow pattern. The typical values of ventilation effectiveness reported in
CEN Report 1752 (CEN, 1998) range between 1.2 and 1.4. If this is the case, the improvement
of the inhaled air quality with PV is likely to be significant, even at lower supply rates of
personalized air.

The mixing of contaminants in the occupied zone may also increase when the personalized
airflow is located in the vicinity of the contaminant source. The results documented that the
use of PV in a room with displacement ventilation increased the concentration of human

57
Chapter 4

bioeffluents and exhaled air in the occupied zone as compared to the case without PV. This
may result in an increased transmission of contaminants between workplaces, and an
increased likelihood of infections for occupants if infectious agents and exhaled air are
associated. Detailed analyses of the risk of infections are performed in Chapter 6. If the
impact of PV on the mixing of contaminants were comparable with the impact of the activity
of occupants, the increase in the exposure of occupants due to the use of PV would be
practically unimportant. Such a comparison, however, remains to be made. The
measurements showed that, although PV promoted mixing, the exposure of the exposed
manikin was always lower than in a room with mixing ventilation.

The ability of PV to increase the transmission of contaminants between workplaces is


associated with the strength and direction of personalized airflow. Because the personalized
airflow from RMP was weaker than the personalized airflow from VDG, the primary cause
for the increased transmission seems to be the airflow direction rather than the strength of
the airflow. Furthermore, the impact of PV on the concentration of contaminants in the
occupied zone may be confounded with the decrease of the stratification height of the
contaminants. It is assumed that the use of PV with an upward airflow direction (VDG)
decreases the stratification height, because the personalized airflow entrains surrounding air
and acts like a thermal plume. The stratification also decreases with the airflow rate from
displacement ventilation (when PV is used). None of these factors (direction and strengths
of PV airflow, height of stratification), however, were studied in detail.

The distribution of human bioeffluents in the room was similar to the distribution of exhaled
air. However, the two contaminants are of a different nature. Human bioeffluents are carried
upward by means of a free convection flow around the body. The airflow of exhalation
penetrates the free convection flow due to its momentum and interacts with the room
airflow. The distributions were similar most probably because of the presence of the heated
office equipment in the workplaces. The thermal flow generated around the equipment
entrained the exhaled air, so that it was carried upward and distributed in the room in the
same way as the bioeffluents.

The use of PV caused large non-uniformities and spatial difference in the distribution of
exhaled air and bioeffluents in front of the workplace where the contaminants were
generated. The distribution of the two contaminants, described in detail in Section 4.6
(reported also in Cermak et al., 2004), was different with the two types of PV. In the rest of
the room the differences between the two contaminants diminished, and the non-
uniformities disappeared. The concentrations of the human-produced contaminants in front
of the polluting workplace were higher than the concentrations in the rest of the room. This
suggests that if there was an occupant seated in front of the polluting workplace, his or her
exposure to the contaminants might have been greater than the exposure of an occupant
seated behind the polluting workplace (i.e. exposed manikin measured in this study).
Therefore, the arrangement of workplaces may be important for the inhaled air quality of
occupants. Because different occupants may produce contaminants (e.g. virulent agents) at
one and the same time, it is difficult to recommend an office layout that would ensure a low
exposure of all occupants at any time. A barrier placed in front of a desk (e.g. a partition)
should in principle prevent the spread of contaminants into the room. This, however,
remains to be studied.

A weak personalized airflow (at a rate of 7 L/s) supplied at the workplace of the exposed
manikin was not able to affect the room distribution of contaminants generated at the
workplace of the polluting manikin. At high flow rates, however, personalized airflow may

58
Chapter 4

cause mixing. At a rate of 15 L/s, the impact of PV on the room air distribution was very
localized. The airflow from VDG was able to cause more mixing than the airflow from RMP.
Because the two terminals supplied air in opposite directions and at different velocities, it is
not certain whether the airflow direction or the velocity was the reason. In the opposite side
of the room, the distribution was affected only slightly and no differences were found
between the two terminals. It is likely that a stronger personalized airflow would promote
mixing of all contaminants in the whole occupied zone, as demonstrated in the study on
PEM (described in Section 1.3) by Faulkner et al. (1993).

Not only the level of air pollution but also air temperature and air humidity influences the
perception of air quality. Fang et al. (1998a, b) confirmed earlier studies and showed that air
is perceived more acceptable with decreasing temperature and humidity. Therefore, the
inhaled air temperature was measured in the present study as well. Without PV, the
temperature was 1°C lower with the displacement ventilation than with the mixing
ventilation. This may cause an improvement of the perceived air quality, depending on the
pollution level. The impact will be more pronounced if the air is clean. The use of PV
providing air 6°C colder than the room air caused the inhaled air temperature to decrease by
up to 5°C (with RMP at 15 L/s). This is in agreement with the results by Melikov et al. (2002)
for CMP and MP, described in Section 1.3. The inhaled air temperature with PV and
displacement ventilation was consistently lower than the inhaled air temperature with PV
and mixing ventilation. The reason was a lower temperature in the occupied zone in the case
of displacement ventilation. The ability of PV to decrease in the inhaled air temperature as
compared to the reference inhaled air temperature did not depend on the total-volume
ventilation system. The analyses presented in Section 4.5 showed that the inhaled air
concentrations are very well correlated with the inhaled air temperatures. The correlation
applies, however, only on the contaminants distributed uniformly (see Section 4.5 for more
details). The correlation allows for a simplification of the perceived air quality analyses as
well as measurements, because information about only one quantity is needed.

The inhaled air quality is assumed to be the most important criterion for the performance
assessment of PV systems. However, the ability of PV to provide cooling at high
temperatures and not to affect thermal sensation at comfortable temperatures is also
important. The present results showed that the whole-body cooling ability of the two types
of PV was about the same. At a rate of 15 L/s, the cooling was equivalent to decreasing the
room air temperature by 1-2°C. The values varied depending on the actual combination of
PV, a total-volume ventilation principle and a thermal manikin (workplace). Such cooling,
even though small, may be sufficient for many occupants who need only minor adjustment
of the local thermal environment.

The whole-body cooling ability of PV is influenced by the strength and direction of


personalized airflow. The air terminals tested in the present study were adjusted according
to people’s preferences identified by Kaczmarczyk et al. (2002b), which may not correspond
to the adjustment providing the greatest cooling. The cooling ability of VDG agreed with the
maximum cooling ability of Climadesk (similar terminal, described in Section 1.3) of 1°C,
reported by Tsuzuki et al. (1999). The maximum cooling ability of another terminal tested by
Tsuzuki et al. (1999), Personal Environment Module (PEM), was equivalent to decreasing the
room temperature by 7°C. The airflow from PEM was, however, much stronger than
airflows examined in the present study. The measurements by Melikov et al. (2002) showed
that the ventilation performance of a strong airflow from PEM is very mediocre (see below).

59
Chapter 4

The thermal comfort of occupants may be affected by the distribution of cooling over the
body. The measurements showed that the cooling impact of both PV types was very
localized (Forejt et al., 2004). The body parts most cooled were those exposed direct to
personalized air, i.e. the head and the chest. The human response to localized cooling is
difficult to predict, because the sensitivity of body parts to air movement, clothing, activity
and preferences of occupants differs in rooms in practice. Nevertheless, the distribution of
cooling provided by PV may not be important so long as the occupants can change the
strength and direction of personalized airflow. Unfortunately, the distribution of cooling
preferred by an occupant may not always ensure an optimum quality of inhaled air.
Therefore, it may be necessary for an occupant to prioritize between a high air quality and
preferred thermal comfort.

The thermal environment provided with both mixing ventilation alone and displacement
ventilation alone was comfortable according to the present standards and guidelines (ISO,
1994; CEN, 1998). The temperatures in the occupied zone were lower with displacement
ventilation than with mixing ventilation due to thermal stratification. The combination of PV
and displacement ventilation increased the air temperature in the occupied zone by approx.
1°C. The measurements with the thermal manikins showed that the changes in the ambient
air temperature did not affect the heat loss from the manikins considerably. This indicates
that the use of PV by some occupants will not stimulate other occupants to adjust their PV.
This would cause a chain reaction with subsequent and continuous changes in thermal
environment, leading to disturbances of occupants and a loss of productivity.

The temperature in the occupied zone increased because of two reasons. The measurements
of velocity identified that PV used at a rate of 15 L/s did not cause mixing of personalized
air and ambient air in the lower occupied zone. Because the airflow rate of displacement
ventilation decreased when PV was used, a smaller volume of conditioned air was involved
in the removal of heat from the lower occupied zone. This caused the increase in
temperature near the floor. The second reason was the ability of PV to increase mixing in the
upper levels of the room (warm air was brought to the occupied zone). This was the case of
VDG, where the ability to cause mixing was greater than that of RMP (at a supply rate of 15
L/s). Therefore, VDG caused the temperature in the occupied zone to increase more than
RMP (Figure 4.19). The consequence of the increase in temperature could be an increase in
the energy consumption for additional conditioning of air, if the temperature in the
occupied zone should be maintained constant.

The normalized (dimensionless) temperature near the floor was 0.4 with displacement
ventilation alone, and it increased to 0.5 when PV was used (Figure 4.19). The value of 0.4
agrees very well with the model by Mundt (1990). According to the model, the temperature
near the floor rises due to the radiation between the ceiling and the walls to the floor and the
following convective transport to the supply air in the floor area. Because the use of PV did
not increase the temperature near the floor substantially, the model by Mundt (1990) can
also be used for rooms with PV and displacement ventilation when a rough estimate is
needed. One should consider that the predicted temperature might underestimate the actual
temperature with the combined system. More measurements at different airflow rates are
needed in order to adjust the model by Mundt (1990) for the combined system. An increase
in the convective heat transfer coefficient might be an option. A coefficient of 6 W/m2
(instead of 3 or 5 W/m2 as used in the original model) provided a better estimate in the
present study.

60
Chapter 4

In the present study, the profiles were not linear (steeper in the lower part of the room,
Figure 4.19). The shape of the profiles reflected the fact that most heat sources were located
in the lower part of the room. In practice, a linear profile is usually used as a first
approximation of the temperature distribution. The analyses of the present results indicate
that the linear profile would underestimate the actual vertical air temperature difference
between head and ankles by roughly 50% (the actual temperature difference was larger),
with or without PV. The assumption of the linear profile is thus questionable in this case.

4.11 Conclusions

• PV will always be able to protect occupants from pollution and thus increase the quality
of inhaled air in rooms with mixing air distribution. The type of a contaminant source
(active or passive, localized or plane) and its location is unimportant. The positioning of
an air terminal device in respect to the occupant and it design influences the quality of
inhaled air to a great extent.

• In rooms with displacement ventilation, the PV will improve the inhaled air quality in
regard to a passive and plane contaminant located on the floor, such as carpet, PVC or
linoleum. The use of PV may, however, increase mixing of contaminants located in the
vicinity of the personalized airflow, such as exhaled air and bioeffluents. This may
increase the transmission of contaminants between workplaces as compared to the case
without PV.

• The inhaled air temperature is well correlated with the inhaled air concentration of
contaminants distributed uniformly. The temperature and concentration are not
correlated if the distribution of the contaminant is non-uniform.

• The impact of a room air distribution principle on the thermal comfort of occupants is
small and insignificant compared to the impact of PV when its maximum cooling
capacity is used. The cooling of the human body provided by PV is rather independent
of the room air distribution generated by a total-volume ventilation system.

• The direction and rate of personalized airflow determine the distribution of active
contaminants generated at workplaces in rooms with displacement ventilation. The use
of PV creates large non-uniformities and spatial difference in the distribution of human-
produced contaminants in the vicinity of a workplace. The distributions of exhaled air
and bioeffluents may be different, depending on the flow direction of personalized air.

• In rooms with displacement ventilation, PV does not affect the distribution of active
contaminants generated in another workplace when the supply airflow rate is low. At a
high rate the type of supply air terminal and its airflow direction may be important.

• The impact of PV on the distribution of temperature and velocity is very localized. PV


does not contribute to the removal of heat from the lower occupied zone in rooms with
displacement ventilation. This increases temperature and consequently the cooling
requirements if the thermal environment should be maintained constant.

61
5
5. Personalized and underfloor ventilation

5.1 Experiment 1

5.1.1 Objectives
The objective of the work presented in this chapter is to investigate the performance of PV in
combination with underfloor air distribution (UFAD).

5.1.2 Experimental conditions


Four passive swirl diffusers (Section 3.3.2) were tested in the present study. A swirl element
mounted in each diffuser allowed for adjustment of air discharge to either a horizontal or a
vertical direction. Both the discharge directions were studied. The experimental design in
terms of an office layout, furniture, manikins, contaminant sources, positions and methods
of measurement were identical with the experiments with mixing and displacement
ventilation, presented in Chapter 4.

The supply air temperatures of both PV and UFAD were 20°C. The total airflow rate
supplied was 80 L/s. These conditions were identical in all experiments. The UFAD system
was controlled according to a VAV strategy, i.e. the amount of air supplied from the floor
decreased proportionally to the increase in the personalized airflow. The conditions tested
(Table 5.1) were designed to allow for a comparison between the mixing ventilation,
displacement ventilation and underfloor ventilation.

Table 5.1. Combinations of personalized airflow rate (L/s) tested.


PV type UFAD vertical discharge UFAD horizontal discharge
Polluting Exposed Polluting Exposed
manikin manikin manikin manikin
- 0 0 0 0
RMP 0 15 0 15
RMP 15 0 15 0
RMP 15 15 15 15
VDG 0 15 0 15
VDG 15 0 15 0
VDG 15 15 15 15

5.1.3 Aerodynamic data of floor diffusers


The evaluations of the throw height and the near zone of the diffusers were based on a
distribution of velocity. The throw height and the near zone were defined as respectively a
horizontal distance and a vertical distance from the centre of the diffuser where the supply
air stream velocity decreases to 0.25 m/s. The supply airflow rates examined were 20 L/s
(case 1, Table 5.1) and 12.5 L/s (cases 4 and 7, Table 5.1) per diffuser. The supply air

63
Chapter 5

temperature was 20°C in both cases. The former condition corresponds to the reference case
without PV (80 L/s through UFAD). The latter condition represents the discharge pattern
when both PV units were used (50 L/s through UFAD + 2 times 15 L/s from PV); however,
PV was not used during the velocity measurements in order not to affect the flow field.

Figure 5.1. Mean air velocity contours for the vertical discharge. Supply airflow rate 20 L/s per
diffuser (left) and 12.5 L/s per diffuser (right).

Figure 5.2. Mean air velocity contours for the horizontal discharge. Supply airflow rate 20 L/s per
diffuser (top) and 12.5 L/s per diffuser (bottom).

Figure 5.1 shows the mean air velocity contours of the airflow provided with the vertical
discharge of UFAD. Only one half of the flow field is presented with the centre of the
diffuser positioned in the origin of the plot. The figure shows that the air spreads in a V-
shape pattern. As expected, the flow dissolved at a lower height when the supply airflow

64
Chapter 5

was weaker. The throw was about 1.0 m and 0.3 m for an airflow rate of 20 and 12.5 L/s per
diffuser, respectively. The contours show that the radius of the near zone was about 0.5 m
and 0.2 m for the two airflow rates.

The velocity distribution with the horizontal discharge is presented in Figure 5.2. The throw
was limited to a very short height, regardless of the airflow conditions. On the other hand,
the near zone radius was much longer compared to the vertical discharge: 0.5 m and 0.2 m at
airflow rates of respectively 20 and 12.5 L/s per diffuser. The area of velocities above 0.1
m/s (where there is a risk of draught) was greater than 1 m.

Figure 5.3. Top: airflow rate 80 L/s, vertical discharger (left) and horizontal discharge (right). Bottom:
Airflow rate 50 L/s, vertical discharge (left) and horizontal discharge (right).

Figure 5.3 shows the smoke visualization for the two discharge patterns. The conditions of
the visualization and of the measurement (Figures 5.1 and 5.2) are identical. Each photo was
taken approx. 10 s after introducing the smoke. The white square grid at the back of the
scene had a module of 0.3 m. Similar to the velocity measurement, the visualization reveals
obvious differences between the vertical and the horizontal discharge and the supply air
conditions. In the case of the horizontal discharge a certain amount of mixing was created
above a height of 0.1 m when the supply airflow reached the walls at a distance of 0.85 m
from the centre of the diffuser as well as when the flows from two diffusers collided (not
shown). The reflection from the walls was not apparent from the velocity contours, which
were measured in the direction parallel to the walls. The visualization was influenced by a
systematic error due to the fact that the smoke was essentially warmer than the supply air.
The supply air temperature was estimated to rise by 1°C to 3°C during the short period the
smoke was introduced to the room.

65
Chapter 5

5.1.4 Inhaled air concentration

Exhaled air and bioeffluents


Figure 5.4 presents the inhaled air concentrations of bioeffluents and exhaled air with
different total-volume ventilation systems alone and in combination with PV. Because the
differences between the two concentrations were much smaller than the differences between
the conditions tested, the concentrations of the bioeffluents and exhaled air were averaged.
The results for mixing ventilation and displacement ventilation (as discussed in Chapter 4)
are presented for comparison.

1
Concentration (CI-CS/CE-CS)

Mixing
0.8 UFAD vertical
UFAD horiz.
0.6 Displacement

0.4

0.2

0
no PV RMP (1) RMP (2) VDG (1) VDG (2)

Figure 5.4. Concentration of exhaled air and bioeffluents (averaged) in the air inhaled by the exposed
manikin with RMP terminal and VDG terminal in combination with different total-volume
ventilation principles. The error bars indicate the maximum uncertainty from the two contaminants
with a level of confidence of 95%. Legend: (1) Exposed manikin using PV at 15 L/s while polluting
manikin does not; (2) both manikins using PV at 15 L/s.

The concentration of human-produced contaminants provided with the horizontal discharge


of UFAD alone was comparable to the concentration provided with the displacement
ventilation alone. With the vertical discharge the concentration in the inhaled air increased
due to mixing in the lower occupied zone. The concentrations provided with the vertical and
the horizontal discharge of UFAD alone were 1.6 and 8.2 times lower than in the case of the
mixing ventilation alone. The use of PV caused the differences between the discharge
patterns to disappear. The comparison of the terminals shows that the RMP terminal
achieved a lower inhaled air concentration of the contaminants than the VDG terminal,
regardless of the total-volume ventilation principle and the pattern of use. The performance
of the terminals was influenced by the interaction of airflows in front of a human body
(discussed in detail in Chapter 4.3). The inhaled air concentrations provided with RMP were
between 0.04 and 0.02. This corresponds to a ventilation effectiveness of 25 to 50. With the
VDG, the concentrations of human-produced contaminants were higher due to the
entrainment of contaminated room air in its very turbulent jet. The values of ventilation
effectiveness ranged between 3 and 6. Comparison with the horizontal discharge of UFAD
(ventilation effectiveness of 8.7) shows that the VDG was not able to decrease the inhaled air
concentration below its reference level.

Figure 5.5 examines the transmission of the human-produced contaminants between the
workplaces. The inhaled air concentration to the exposed manikin with underfloor, mixing
and displacement ventilation is compared. The polluting manikin was using PV at 15 L/s
while the PV of the exposed manikin was not used. With the UFAD system, the transmission
increased with both types of air terminal device. The transmission was lower with the VDG

66
Chapter 5

than with the RMP for the combinations tested. The concentration of contaminants was
somewhat lower with the horizontal discharge of UFAD than with the displacement
ventilation.

1
Concentration (CI-CS/CE-CS)

Mixing
0.8 UFAD vertical
*) UFAD horiz.
0.6 Displacement

0.4

0.2

0
no PV RMP VDG

Figure 5.5. Concentration of exhaled air and bioeffluents (averaged) in the air inhaled by the exposed
manikin with PV in combination with different total-volume ventilation principles. Polluting
manikin using PV at 15 L/s while exposed manikin does not. The error bars indicate the maximum
uncertainty from the two contaminants with a level of confidence of 95%. *) There was a difference
between exhaled air and bioeffluents – see Appendix B.

Floor contaminant
The analyses showed that there were no differences in the inhaled air concentration of the
floor contaminant between the four total-volume ventilation systems. No differences were
identified between the two manikins either. The normalized concentration in the inhaled air
of both the manikins did not differ by more than 5-10% from unity. The use of PV did not
have any adverse effects on the inhaled air quality of an unprotected manikin. This was to
be expected, because the concentration of the floor contaminant was largely uniform. Lower
concentrations were recorded with both the UFAD and the displacement ventilation near
the floor. However, the entrainment of room air in the upward boundary layer flow
increased the concentration almost to the room air level before it was inhaled. The use of PV
decreased the inhaled air concentration several times, depending on the personalized air
terminal used. Appendix E presents the details of the floor contaminant concentration in the
inhaled air.

5.1.5 Inhaled air temperature


Figure 5.6 compares the inhaled air temperature of the polluting manikin with the different
ventilation systems tested, with and without PV. The PV airflow rate of the polluting
manikin was 15 L/s, while the exposed manikin did not use PV. The inhaled air
temperatures with both the discharge patterns of UFAD alone were between the
temperatures of the mixing and displacement ventilation. The figure shows that the RMP
provided a lower inhaled air temperature than the VDG. The results are well correlated with
the inhaled air concentration of the floor contaminant (Section 4.5). Regardless of the total-
volume ventilation system, the temperatures were about the same for each terminal type,
except for the RMP combined with the mixing ventilation. The reason can be the airflow
direction of RMP, which was re-adjusted between the experimental stages 1 and 2.

67
Chapter 5

28
27 Mixing
UFAD vertical
26
Temperature (°C)

UFAD horiz.
25 Displacement
24
23
22
21
20
no PV RMP VDG

Figure 5.6. Inhaled air temperature of the polluting manikin with different total-volume ventilation
principles. Polluting manikin using PV at 15 L/s while exposed manikin does not. The error bars
indicate the maximum uncertainty from the two contaminants with a level of confidence of 95%.

5.1.6 Contaminant distribution


The distribution of the floor contaminant was generally uniform under all scenarios tested,
except for near the floor, where non-uniformities were identified due to the close proximity
of the supply air terminals and a coarse grid of the tracer-gas dosing points on the floor. The
distribution was very similar to the distribution with displacement ventilation. Therefore,
only the distribution of the human-produced contaminants is presented in this section.
There were no differences found between the concentrations of bioeffluents and exhaled air
for any combination of PV with underfloor ventilation. The concentrations of the two
contaminants were thus averaged in most of the following figures in order to provide
generalization.

Total-volume ventilation alone


Figure 5.7 compares the distribution of the human-produced contaminants (average of
bioeffluents and exhaled air) with the underfloor, mixing and displacement ventilation
alone. The concentrations recorded in front of the polluting manikin and next to the exposed
manikin were averaged in order to provide a more general picture of the distribution. The
distribution for the horizontal discharge of UFAD was identical with the distribution for the
displacement ventilation. The concentrations are very low near the floor, and they increased
with the room height. The exact horizontal position of the interface layer between the lower
cleaner zone and the upper more contaminated zone is difficult to identify. The
concentrations begin to stratify between the heights of 0.6 and 1.1 m. With the vertical
discharge the concentrations in the occupied zone were higher compared to the horizontal
discharge. The vertical swirl flow penetrated the upper contaminated zone and brought
contaminants to the occupied zone. The interface layer stretched between the heights of 1.1
and 1.7 m. With all the systems, the concentrations exceeded a value of 1 above the occupied
zone (i.e. exhaust air concentration). Analyses showed that the profiles measured in front of
the polluting manikin overestimated the average contaminant distribution in the room,
apparently because of the proximity of the source.

68
Chapter 5

2.5
Height above floor (m)

1.5
Mixing
1 UFAD vertical
UFAD horizontal
0.5 Displacement

0
0 0.5 1 1.5 2
Concentration (C-CS/CE-CS)

Figure 5.7. Concentration profiles of human bioeffluents and exhaled air (averaged) near the two
workplaces (positions A and B, averaged) with total-volume ventilation systems alone.

Personalized airflow close to a source


The distribution of the contaminant varied considerably depending on the relation between
personalized airflow and the location of a contaminant source. The situation was similar to
the combination of PV and displacement ventilation. Figure 5.8 presents separately the
concentration of human bioeffluents and exhaled air measured in front of the polluting
manikin (position A). The polluting manikin was using PV at 15 L/s, while the exposed
manikin did not use PV. Figure 5.9 presents the concentration of the human-produced
contaminants measured next to the exposed manikin (position B) under the same condition.
The use of PV increased the concentration of both the human-produced contaminants in the
occupied zone. In all cases the VDG terminal achieved lower concentration levels in the
occupied zone. The concentrations were higher with the vertical discharge than the
horizontal discharge. The distributions with the vertical discharge were similar for the two
positions, while with the horizontal discharge they were not similar. Less mixing in the
lower occupied zone with the horizontal discharge allowed for stratification of the
contaminants in a horizontal layer at a height of 0.6 m. Although lower in magnitude, a
similar concentration peak was observed also in experiments with displacement ventilation.

Personalized airflow far from a source


Figure 5.10 documents the impact of the RMP on the distribution of human bioeffluents and
exhaled air that was generated at another workplace. The exposed manikin used PV and the
polluting manikin did not. The profiles recorded near the two manikins (positions A and B)
were averaged. The concentration in the occupied zone decreased when the RMP was used,
most probably as a result of decreasing the throw of UFAD. The distribution of the
contaminants was not affected with the combination of the RMP and the horizontal
discharge. The concentration remained very low near the floor. Figure 5.11 shows the
distribution for the VDG terminal. When the horizontal discharge was employed, the mixing
introduced by the VDG caused an increase of the concentrations near the floor (compared
with the RMP that did not change the concentration). At the same time, the concentration
decreased near the ceiling. With the VDG and the vertical discharge the concentrations
decreased near the floor. It is impossible, however, to conclude with any certainty, whether
the decrease in the throw or the use of PV was the cause.

69
Chapter 5

2.5
Vertical alone
Height above floor (m)

2 Vert. + VDG
Vert. + RMP
1.5 Horizontal alone
Hor. + VDG
1 Hor. + RMP

0.5

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Bioeffluents (C-CS/CE-CS)

2.5
Vertical alone
Height above floor (m)

2 Vert. + VDG
Vert. + RMP
1.5 Horizontal alone
Hor. + VDG
1 Hor. + RMP

0.5

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Exhaled air (C-CS/CE-CS)
Figure 5.8. Concentration profiles of bioeffluents (top) and exhaled air (both) in front of the polluting
manikin (position A) with PV and two discharge patterns of UFAD. Polluting manikin using PV at
15 L/s while exposed manikin does not.

2.5 2.5 Vertical alone


Vert. + VDG
Height above floor (m)
Height above floor (m)

2 2 Vert. + RMP
Horizontal alone
1.5 1.5
Hor. + VDG
Hor. + RMP
1 1

0.5 0.5

0 0
0 0.5 1 1.5 2 0 0.5 1 1.5 2
Concentration (C-CS/CE-CS) Concentration (C-CS/CE-CS)
Figure 5.9. Concentration profiles of exhaled air and bioeffluents (averaged) next to the exposed
manikin (position B) with PV and two discharge patterns of UFAD. Polluting manikin using PV at
15 L/s while exposed manikin does not.

70
Chapter 5

2.5
Height above floor (m)

1.5

1 Vertical alone
Horizontal alone
0.5 RMP + vertical
RMP + horizontall
0
0 0.5 1 1.5 2
Concentration (C-CS/CE-CS)

Figure 5.10. Concentration profiles of exhaled air and bioeffluents (averaged) near the two workplaces
(positions A and B, averaged) with RMP and both discharge patterns of UFAD. Exposed manikin
using PV at 15 L/s while polluting manikin does not.

2.5
Height above floor (m)

1.5

1 Vertical alone
Horizontal alone
0.5 VDG + vertical
VDG + horizontal
0
0 0.5 1 1.5 2
Concentration (C-CS/CE-CS)

Figure 5.11. Concentration profiles of exhaled air and bioeffluents (averaged) near the two workplaces
(positions A and B, averaged) with VDG and both discharge patterns of UFAD. Exposed manikin
using PV at 15 L/s while polluting manikin does not.

5.1.7 Temperature distribution


The temperature distribution was assessed using the vertical air temperature profiles
measured in front of the polluting manikin and next to the exposed manikin. Analyses
showed that the temperature profiles were very similar at the two positions. Figure 5.12
compares the normalized temperature profiles of underfloor ventilation with the profiles
measured for mixing and displacement ventilation. Personalized ventilation was not used in
any of the cases presented. It is shown that the temperatures near the floor were equal to
approx. 0.6 and 0.7 with the horizontal and the vertical discharge of underfloor ventilation,
respectively. The normalized temperature with the displacement ventilation was 0.4, as
discussed in Chapter 4. However, already at a height of 0.6 m the differences between the
discharge patterns had vanished and the distribution with UFAD became similar to the
distribution with displacement ventilation.

71
Chapter 5

2.5
Displacement
Height above floor (m)

2 UFAD vertical
UFAD horizontal
1.5 Mixing

0.5

0
0.2 0.4 0.6 0.8 1 1.2
Temperature (T-TS/TE-TS)

Figure 5.12. Mean air temperature profiles next to the two workplaces (positions A and B, averaged)
with different ventilation principles without PV.

Figure 5.13 documents the impact of PV and underfloor ventilation on the temperature
distribution when both PV units supplied air at 15 L/s. The use of the VDG increased the
temperature in the occupied zone compared to the reference case by 0.15 (0.9°C) and 0.25
(1.5°C) when it was combined respectively with the vertical and the horizontal discharge.
The RMP combined with the horizontal discharge increased the temperature by as little as
0.1 (0.5°C) and it did not increase the temperature to any degree when combined with the
vertical discharge. The increase of temperature was pronounced most at a height of 0.6 m.
Above a height of 1.4 m the temperatures were generally unaffected. Air temperature
decreased below the ceiling when the VDG terminal was used. The analysis of the
individual profiles identifies that this was the case only near the exposed manikin, where the
upward airflow from the VDG had a direct influence on the temperature measurement. The
same effect was identified also with the displacement ventilation (Figure 4.19).

2.5 2.5
Vertical alone Horizontal alone
Height above floor (m)

Height above floor (m)

2 Vertical + RMP 2 Horizontal + RMP


Vertical + VDG Horizontal + VDG
1.5 1.5

1 1

0.5 0.5

0 0
0.4 0.6 0.8 1 1.2 0.4 0.6 0.8 1 1.2
Temperature (T-TS/TE-TS) Temperature (T-TS/TE-TS)

Figure 5.13. Mean air temperature profiles next to the two workplaces (positions A and B, averaged)
with the vertical discharge (left) and the horizontal discharge (right). Both manikins using PV at 15
L/s.

5.1.8 Velocity distribution


Figure 5.14 presents the average profiles of mean velocity at the two locations with the four
total-volume ventilation systems tested. The profiles measured near the floor for the
displacement ventilation and the horizontal discharge of UFAD are similar due to the air
current that spreads in a thin layer along the floor. The highest velocity of around 0.16 m/s
was measured at a height of 0.05 m, and it decreased rapidly to a level of 0.05 m/s already at

72
Chapter 5

a height of 0.2 m. The velocities are lower with the vertical discharge of UFAD. Above a
height of 1 m the differences between the displacement ventilation and both the discharge
patterns of UFAD vanished. The highest velocities among the tested systems were identified
for the mixing ventilation. Section 4.9 offers an explanation based on the airflow pattern.

2.5
Displacement
Height above floor (m)

2 UFAD vertical
UFAD horizontal
1.5 Mixing

0.5

0
0 0.05 0.1 0.15 0.2
Velocity (m/s)

Figure 5.14. Mean air velocity profiles next to the two workplaces (positions A and B, averaged) with
different total-volume ventilation principles without PV.

2.5
Vert. alone
Height above floor (m)

2 Vert. + RMP
Vert. + VDG
1.5 Hor. alone
Hor. + RMP
1 Hor. + VDG

0.5

0
0 0.05 0.1 0.15 0.2
Velocity (m/s)

Figure 5.15. Mean air velocity profiles next to the two workplaces (positions A and B, averaged) with
PV and underfloor ventilation. Both manikins using PV at 15 L/s.

Figure 5.15 presents the velocity distribution with the UFAD system alone and in
combination with PV. The previously high velocity near the floor decreased when PV was
used due to the decrease in the underfloor ventilation rate. The VDG terminal is shown to
increase the velocities in the upper zone, regardless of the discharge direction of UFAD (also
identified with the displacement ventilation, Figure 4.20). The velocities did not exceed a
level of 0.05 m/s above a height of 0.2 m (incl.) with the RMP or any of the discharge
patterns of UFAD.

73
Chapter 5

5.2 Experiment 2
The results presented in the previous chapter indicated that the throw height of UFAD had
an impact on the distribution of contaminants, temperature and velocity in the room. In
order to ensure a comfortable thermal environment, the volume of air supplied from the
floor decreased proportionally with an increase in the PV airflow rate. The basic problem
was that the operation of the two systems was related, and hence their impact on the
distribution confounded with each other, which made the evaluation of the impact of the
throw height difficult.

5.2.1 Objectives
The objective of this part of the study is to identify the impact of the throw height of UFAD
on the performance of PV in rooms with UFAD.

5.2.2 Experimental conditions


Two levels of the throw provided with a vertical discharge of UFAD were investigated.
Unlike in the previous chapter, the throw was maintained constant regardless of the use of
PV. This was achieved by supplying a constant amount of air from the floor. In order to
ensure a comfortable thermal environment, however, the temperature of air supplied from
the floor had to be increased with an increase in the PV airflow rate. The changes in the
supply air temperature did not have a large impact on the throw height (see below). The
supply airflow rate of the floor diffusers was 80 L/s (identical with the previous
experiments) and 50 L/s. The throw heights, as reported in Section 5.1.3, were 1.0 m and 0.3
m respectively in the two cases. Both RMP and VDG were tested at a supply rate of 15 L/s.
The combinations of PV airflow rates tested (patterns of PV use) were identical with those
tested in the previous stage (see Table 5.1, substitute the vertical discharge and the
horizontal discharge with respectively a long throw and a short throw).

Table 5.2. Combinations of airflow rates and temperatures of PV and UFAD.


PV supply UFAD supply Exhaust
Airflow rate Temperature Airflow rate Temperature Airflow rate Temperature
(L/s) (°C) (L/s) (°C) (L/s) (°C)
0 - 80 20 80 26
15 20 80 21.1 95 26
15 + 15 20 80 22.2 110 26
0 - 50 16.4 50 26
15 20 50 18.2 65 26
15 + 15 20 50 20 80 26

Table 5.2 presents the combinations of the supply air volume and temperature of
personalized and underfloor ventilation systems arising from the constant air volume
control strategy applied. The UFAD system was not under thermostatic control. The supply
air temperature was determined from a calculation in order to provide the same cooling
capacity of the supply air. The impact of the supply air temperature on the vertical throw
was examined at a rate of 50 L/s (12.5 L/s per diffuser). Figure 5.16 shows the mean air
velocity contours at a supply air temperature of 20°C and 16.4°C. The centre of the diffuser
is located in the origin of the plot. It is shown that the supply air temperature did not have a

74
Chapter 5

substantial impact on the throw height. It should be noted, however, that the room air
temperature was approx. 1°C higher when the supply air temperature was 20°C than when
it was 16.4°C. The reason was a lower cooling ability of the supply air, because PV was not
used in either of the cases in order to avoid its impact on the distribution.

Figure 5.16. Mean air velocity contours for the vertical discharge at a supply airflow rate 12.5 L/s per
diffuser. Supply air temperature 20°C (left) and 16.4°C (right).

A consistent similarity between the distribution of bioeffluents and exhaled air was
identified in most cases presented in the previous chapter. The experimental design was
modified in that the tracer-gas used to simulate the bioeffluents of the polluting manikin
was applied to mark the air exhaled from the exposed manikin. Human bioeffluents were
thus not simulated in this part of the study. Therefore, because both manikins acted as a
source of contaminants, the manikins referred previously to as the polluting manikin and
the exposed manikin are referred to in the following as respectively the front manikin and
the back manikin. The experimental design in terms of an office layout, furniture, positions
and methods of measurement was not changed.

5.2.3 Inhaled air concentration


The inhaled air concentrations of the floor contaminant and exhaled air are presented in
detail in Appendix E. Figure 5.17 shows the concentration of exhaled air produced by one
manikin in the air inhaled by the other manikin at the two throw heights of UFAD. The
manikin presented was using PV at 15 L/s, while the other (polluting) manikin did not use
PV. It is shown that the throw height of UFAD did not have an impact on the inhaled air
concentration, although different levels of concentration were provided with the two
terminals. RMP is shown to achieve a very low contaminant concentration (corresponding to
a ventilation effectiveness of 50). The concentration provided with VDG was higher due to
its low efficiency.

75
Chapter 5

Concentration (C-CS/CE-CS) 1 1

Concentration (C-CS/CE-CS)
UFAD
0.8 0.8 throw height
1.0 m (80 L/s)
0.6 0.6 0.3 m (50 L/s)

0.4 0.4

0.2 0.2

0 0
RMP VDG no PV RMP VDG no PV

Figure 5.17. Concentration of exhaled air in the air inhaled by the front manikin (left) and the back
manikin (right). The manikin presented is using PV at 15 L/s, while the other manikin (producing
contaminants) does not.

The only drawback of PV identified in the previous chapters was associated with the ability
of PV to increase mixing of indoor air, and thus to increase the transport of human-
produced contaminants between occupants. Figure 5.18 compares the inhaled air
concentration of exhaled air at two different throw heights of UFAD. The exposure of both
the manikins to a contaminant generated by the other manikin is presented. The manikin
presented did not use its PV while the other (polluting) manikin used PV at 15 L/s. It is
shown that the throw height of UFAD affects the transmission of contaminants between
occupants. With the long throw there was almost no difference between the two terminals,
while the VDG terminal performed better than the RMP terminal when the throw was short.
The results were very similar for the two manikins. This means that the transmission of
exhaled air from the front manikin to the back manikin is comparable with the transmission
of exhaled air from the back manikin to the front manikin.

1 1
Concentration (C-CS/CE-CS)

Concentration (C-CS/CE-CS)

UFAD
0.8 0.8 throw height
1.0 m (80 L/s)
0.6 0.6 0.3 m (50 L/s)

0.4 0.4

0.2 0.2

0 0
RMP VDG no PV RMP VDG no PV

Figure 5.18. Concentration of exhaled air in the air inhaled by the front manikin (left) and the back
manikin (right). The manikin presented does not use PV, while the other manikin (producing
contaminants) uses PV at 15 L/s.

5.2.4 Inhaled air temperature


The inhaled air temperatures of the manikins were identical with the two throw heights,
although the supply air temperature UFAD was different for the two discharge patterns
(Table 5.2). Analyses showed that neither the supply conditions of UFAD nor the use of PV
by the other manikin were able to affect the inhaled air temperatures substantially. The use
of PV decreased the inhaled air temperature of a manikin exposed to PV. The temperatures
were comparable with the temperatures measured in the previous chapter. No impact of the
throw height or temperature of UFAD on the inhaled air temperature was found.

76
Chapter 5

5.2.5 Contaminant distribution


The distribution of the floor contaminant was similar to the distribution of the floor
contaminant identified in the previous section, and hence similar to the distribution with the
displacement ventilation. Therefore, only the distribution of exhaled air is reported in the
following. Both manikins acting as a source of exhaled air made it possible to use two
experiments to evaluate the distribution of the exhaled air under an analogous setting but in
an opposite layout. For example, the distribution of the contaminant exhaled from the front
manikin using PV while the back manikin did not use PV was supplemented with the
distribution of the contaminants exhaled from the back manikin using PV while the front
manikin did not use PV. The contaminant profiles measured next to the two manikins in the
two cases were similar and therefore averaged in order to provide a more general picture of
the distribution.

Personalized airflow close to a source


Figures 5.19 and 5.20 present the vertical concentration profiles of exhaled air for UFAD
alone and in combination with respectively the RMP and the VDG terminal. The manikin
generating contaminants was using PV at 15 L/s, while the other manikin did not use PV.
The contaminant distribution was clearly stratified with the UFAD alone. When the throw
was short, the concentrations were very low in the lower occupied zone. A long throw
obviously affected the contaminants collected above the occupied zone and brought them to
the lower zone. The concentration near the floor increased to about 50% of the concentration
in the exhaust. The use of the RMP promoted diffusion of the exhaled air in the room air and
thus destroyed the stratification (Figure 5.19). The concentration profiles did not depend on
the throw height. The use of VDG also increased the concentration in the occupied zone.
However, the stratification was not destroyed completely. Lower concentrations were
measured near the floor for the combination of the VDG and the short throw, as compared
to the long throw.

Figure 5.21 presents the vertical contaminant profiles of exhaled air with the VDG terminal
combined with UFAD. Both VDG units were used at 15 L/s. The throw height of UFAD was
again constant, whether or not PV was used. The use of both the VDG units achieved a
uniform distribution when the throw of UFAD was long; however, the stratification
decreased but remained when the throw was short. Comparison of Figures 5.20 and 5.21
shows that the use of both VDG units at a rate of 15 L/s increased the concentration in the
occupied zone as compared to the case when only one VDG unit was used.

2.5
UFAD alone
throw 0.3 m (50 L/s)
Height above floor (m)

2
UFAD alone
throw 1.0 m (80 L/s)
1.5
UFAD + RMP
1 throw 0.3 m (50 L/s)
UFAD + RMP
0.5 throw 1.0 m (80 L/s)

0
0 0.5 1 1.5 2
Concentration (C-CS/CE-CS)

Figure 5.19. Concentration profiles of exhaled air. RMP terminal directed towards the source in
combination with the vertical discharge of UFAD. Personalized airflow rate 15 L/s.

77
Chapter 5

2.5
UFAD alone
throw 0.3 m (50 L/s)
Height above floor (m)

2
UFAD alone
throw 1.0 m (80 L/s)
1.5
UFAD + VDG
1 throw 0.3 m (50 L/s)
UFAD + VDG
0.5 throw 1.0 m (80 L/s)

0
0 0.5 1 1.5 2
Concentration (C-CS/CE-CS)

Figure 5.20. Concentration profiles of exhaled air. VDG terminal directed towards the source in
combination with the vertical discharge of UFAD. Personalized airflow rate 15 L/s.

2.5
UFAD alone
throw 0.3 m (50 L/s)
Height above floor (m)

2
UFAD alone
throw 1.0 m (80 L/s)
1.5
UFAD + VDG
1 throw 0.3 m (50 L/s)
UFAD + VDG
0.5 throw 1.0 m (80 L/s)

0
0 0.5 1 1.5 2
Concentration (C-CS/CE-CS)

Figure 5.21. Concentration profiles of exhaled air. VDG terminal directed towards the source in
combination with the vertical discharge of UFAD. Both VDG units used at 15 L/s.

Personalized airflow far from a source


Figure 5.22 presents the vertical concentration profiles of exhaled air with the RMP and
UFAD. The manikin generating contaminants did not use PV, while the other manikin used
PV at 15 L/s. It is shown that the use of the RMP (not directed towards the source) did not
affect the contaminant distribution in the occupied zone substantially at any throw height of
UFAD. The differences between the contaminant profiles are caused due to small local non-
uniformities. The exhaled air concentration profiles of the VDG and UFAD are shown in
Figure 5.23. The conditions are identical with the previous figure for RMP. Unlike the RMP
terminal, the VDG terminal created mixing both in the occupied zone and below the ceiling
in comparison with the UFAD system alone. The concentration in the lower occupied zone
increased in the case of the short throw, but it did not change from the reference when the
throw was long. This indicates that the long throw dominated the airflow pattern in the
lower occupied zone.

78
Chapter 5

2.5
UFAD alone
throw 0.3 m (50 L/s)
Height above floor (m)

2
UFAD alone
throw 1.0 m (80 L/s)
1.5
UFAD + RMP
1 throw 0.3 m (50 L/s)
UFAD + RMP
0.5 throw 1.0 m (80 L/s)

0
0 0.5 1 1.5 2
Concentration (C-CS/CE-CS)

Figure 5.22. Concentration profiles of exhaled air. RMP terminal not directed towards the source in
combination with the vertical discharge of UFAD. Personalized airflow rate 15 L/s.

2.5
UFAD alone
throw 0.3 m (50 L/s)
Height above floor (m)

2
UFAD alone
throw 1.0 m (80 L/s)
1.5
UFAD + VDG
1 throw 0.3 m (50 L/s)
UFAD + VDG
0.5 throw 1.0 m (80 L/s)

0
0 0.5 1 1.5 2
Concentration (C-CS/CE-CS)

Figure 5.23. Concentration profiles of exhaled air. VDG terminal not directed towards the source in
combination with the vertical discharge of UFAD. Personalized airflow rate 15 L/s.

5.2.6 Temperature distribution


The temperature in the occupied zone was affected by the fact that the supply air
temperature of UFAD was varied in order to provide a constant cooling ability for the room
(Table 5.2). As in the previous chapter, the temperature distribution was assessed based on
the vertical profiles of the mean air temperature recorded near the two workplaces. The
profiles from the two positions were averaged because they were similar. Figure 5.24
presents the dimensionless temperature profiles with PV and UFAD for the two different
throw heights. Both PV units were used at 15 L/s. The situation represents the case when
the impact of PV was the greatest. The comparison of the reference cases showed that the
larger throw height increased mixing in the occupied zone and provided higher
temperatures near the floor compared to the lower throw height. The temperature profile
was steeper (larger stratification) when the throw was lower. The use of PV increased the
temperature in the occupied zone due to the increase in the airflow rate and temperature of
UFAD as well as to the mixing generated by PV. With the RMP the temperatures did not
increase significantly. The largest increase in the normalized temperature with RMP was 0.1
when combined with the vertical discharge providing a throw of 1.0 m (Figure 5.24, left).
The temperatures increased more with the VDG than with the RMP. In the case of the long
throw the normalized temperature increased by as much as 0.2, so that the stratification was
almost destroyed. The differences between the air terminals and the throw heights vanished
above a height of 1.4 m. The temperature below the ceiling decreased with the VDG

79
Chapter 5

regardless of a combination of the discharge direction and the airflow rate. The same
characteristic, caused by the upward direction of personalized airflow, was identified also in
the previous section.

Vertical discharge - 80 L/s Vertical discharge - 50 L/s


2.5 2.5
Height above floor (m)

Height above floor (m)


2 no PV 2 no PV
RMP RMP
1.5 VDG 1.5 VDG

1 1

0.5 0.5

0 0
0.4 0.6 0.8 1.0 1.2 0.4 0.6 0.8 1.0 1.2
Temperature (T-TS/TE-TS) Temperature (T-TS/TE-TS)

Figure 5.24. Mean air temperature profiles next to the two workplaces (positions A and B, averaged)
with the vertical discharge of UFAD providing a throw of 1.0 m (left) and a throw of 0.3 m (right).
Both manikins using PV at 15 L/s.

5.2.7 Velocity distribution


The distribution of velocity is not reported in detail because it was similar to the distribution
with the vertical discharge identified in the previous experiment. The highest velocities were
again measured in the lower occupied zone; however, they did not exceed a value of 0.07
m/s at any height.

80
Chapter 5

5.3 Experiment 3
The results presented in the previous chapters showed that PV supplying 15 L/s towards a
contaminant source was able to promote mixing of the contaminant in the occupied zone
and thus increase the transmission of the contaminant between workplaces.

5.3.1 Objectives
The objective of this section is to investigate the impact of the strength of PV airflow on
mixing in the occupied zone.

5.3.2 Experimental conditions


The ability of PV to decrease mixing in the occupied zone was studied with both RMP and
VDG supplying air at a rate of 7 L/s. The experiments are summarized in Table 5.3.
Although different discharge patterns were used (experiments were not performed in the
same period of time), they both provided a short throw and were thus not supposed to affect
mixing in the occupied zone significantly. The ability of RMP to increase mixing of the
contaminants generated at another workplace at a high supply rate of 30 L/s was tested as
well. The supply air temperature of PV was 20°C in all cases. The experimental design in
terms of an office layout, furniture, positions and methods of measurement was identical
with those of the previous experiments.

Table 5.3. Experimental conditions. Impact of the strength of PV airflow on mixing in the occupied
zone.
PV UFAD Contaminant
Terminal Airflow rate (L/s) Discharge Airflow rate Temperature source
Front Back (L/s) (°C)
manikin manikin
VDG 7 0 vertical 50 17.2 as in 5.2
VDG 0 7 vertical 50 17.2 as in 5.2
RMP 7 0 horizontal 73 20 as in 5.1
RMP 0 30 horizontal 50 20 as in 5.1

5.3.3 Upward airflow direction


Figure 5.25 examines the impact of the VDG airflow rate (7 L/s and 15 L/s) on the
transmission of exhaled air between occupants. Similar results were obtained for the two
manikins. It is shown that the transmission of contaminants increase even at a rate of 7 L/s,
but not as much as at a rate of 15 L/s. Figure 5.26 presents the vertical concentration profiles
of the exhaled air with the VDG under the same conditions. The distribution with the VDG
directed towards the manikin generating contaminants while the other manikin did not use
PV is presented. So as in Experiment 2 (Section 5.2.5), the figure is based on two experiments
with an analogous setting but an opposite layout (Table 5.3). It is shown that the VDG
providing 7 L/s increased the concentration in the occupied zone compared to the reference
case of UFAD alone; the increase was not as large as at a rate of 15 L/s. Although the
concentrations increased, the stratification was preserved in all cases.

81
Chapter 5

Concentration (C-CS/CE-CS) 1

Inhaled to front manikin


0.8 back manikin

0.6

0.4

0.2

0
no PV VDG: 7 L/s VDG: 15 L/s

Figure 5.25. Concentration of exhaled air in the air inhaled by both manikins. VDG combined with a
vertical discharge of UFAD, short throw. The manikin presented did not use PV, while the other
used PV at 7 L/s or 15 L/s.

2.5
UFAD alone
Height above floor (m)

2 UFAD + VDG 7 L/s


UFAD + VDG 15 L/s
1.5

0.5

0
0 0.5 1 1.5 2
Concentration (C-CS/CE-CS)

Figure 5.26. Concentration profiles of exhaled air. VDG terminal directed towards the source in
combination with the vertical discharge of UFAD, short throw. Personalized airflow rate of 7 L/s and
15 L/s.

5.3.4 Horizontal or downward airflow direction


Figure 5.27 documents the impact of RMP airflow rate on the transmission of bioeffluents
and exhaled air from the front manikin to the back manikin. The measurements showed that
the transmission of the contaminants was comparable for the airflow rates of 7 L/s and 15
L/s. Although the airflow did not reach the face of the manikin directly at 7 L/s (see smoke
visualization, Figure 4.1), the mixing was apparently high enough to promote diffusion of
the contaminants to the room. The figure also documents the similarity between the human
bioeffluents and exhaled air. Figure 5.28 compares the distribution of the human-produced
contaminant at the two airflow rates of RMP directed towards the front manikin. The
profiles were very different at the two positions. At a lower rate of 7 L/s the RMP did not
support the stratification of the contaminants in the form of a peak in front of the polluting
manikin. The concentrations were uniform and due to the proximity of the source, higher
than 1. Further from the source, i.e. next to the back manikin, the contaminant profiles were
similar. This reveals the reason for the almost identical inhaled air concentrations presented
in Figure 5.27.

82
Chapter 5

Concentration (C-CS/CE-CS) 1
Bioeffluents
0.8
Exhaled air

0.6

0.4

0.2

0
no PV RMP: 7 L/s RMP: 15 L/s
Figure 5.27. Concentration of bioeffluents and exhaled air in the air inhaled by the exposed manikin
with RMP and horizontal discharge of UFAD. Polluting manikin using PV at 7 L/s and 15 L/s while
back (exposed) manikin does not.

2.5 Horizontal alone, A


Height above floor (m)

RMP 7 L/s, A
2
RMP 15 L/s, A
Horizontal alone, B
1.5
RMP 7 L/s, B
1 RMP 15 L/s, B

0.5

0
0 0.5 1 1.5 2
Concentration (C-CS/CE-CS)

Figure 5.28. Concentration profiles of exhaled air and bioeffluents (averaged) next to the two
workplaces (positions A and B) with RMP and the horizontal discharge of UFAD. Front manikin
using PV at 7 L/s and 15 L/s while back manikin does not.

Figure 5.29 presents the vertical concentration profiles of the human-produced contaminants
produced by the front manikin while the back manikin used PV at a rate of 15 L/s and 30
L/s. The front manikin did not use PV. The distributions measured near the two workplaces
were very similar, despite the different distance from the source and the PV, and are thus
averaged. It is shown that the airflow rate had little impact on the contaminant distribution.

2.5
Height above floor (m)

1.5

1
Horizontal alone
0.5 RMP 15 L/s
RMP 30 L/s
0
0 0.5 1 1.5 2
Concentration (C-CS/CE-CS)

Figure 5.29. Concentration profiles of exhaled air and bioeffluents (averaged) recorded next to the two
workplaces (positions A and B, averaged) with RMP and the horizontal discharge of UFAD. Exposed
manikin using PV at 15 L/s and 30 L/s while front (polluting) manikin does not.

83
Chapter 5

5.4 Discussion
Two types of air terminal device for PV were combined with UFAD discharging air in a
vertical and a horizontal direction. The impact of the vertical throw height of UFAD and the
airflow rate of PV were studied.

Knowledge concerning the distribution of different contaminants in rooms with UFAD


systems (without PV) is limited. The literature (Arens et al., 1991; Bauman et al., 1991; Fisk et
al., 1991; Loudermilk, 1999; REHVA, 2003; Bauman and Daly, 2003) indicates that on the one
hand the performance of UFAD providing a short throw resembles displacement
ventilation. On the other hand, the ventilation performance of UFAD with a long throw was
associated with mixing ventilation. The impact of different throw heights of UFAD on the
inhaled air quality was not, however, studied in detail.

The analyses of the contaminant distribution measured in the present study showed that the
performance of UFAD with a horizontal discharge of air (throw of 0.1 m) was similar with
the performance of displacement ventilation. In both cases the supply air spread along the
floor in a thin layer, which made it possible for the free convection flow around the body to
entrain and transport clean air upward to the breathing zone (performance in regard to
thermal comfort was different, see below). UFAD with a short vertical throw (0.3 m)
provided a similar pattern, although both the distribution of velocity (Figure 5.14) and the
smoke visualization (Figure 5.3) indicated more mixing in the lower occupied zone. As
expected, UFAD with a longer vertical throw penetrated into the occupied zone and brought
polluted air from the higher levels towards the floor. The results showed that the throw
height just comparable to the height of the breathing zone (1.0 m) increased the
concentration of human-produced contaminants in the lower zone to about 50% of the
concentration in the exhaust air. Mixing conditions were not created, apparently because the
throw was still too short.

The results of measurements were used in order to verify two physical models of UFAD
systems suggested recently. Lin and Linden (2002) discussed a situation with one heat
source and one supply opening. However, the model could not be applied in the present,
more complex case due to the large number of simplifications it contained. Yamanaka et al.
(2002) proposed a model that aims at predicting the vertical distribution of human-produced
contaminants. The model was verified in a scaled experimental room. The distribution
predicted for the conditions tested was in a good agreement with the distribution obtained
from the measurement. This is documented in Appendix F. Although more experimental
data are needed for further verification, the model by Yamanaka et al. (2002) could become
an efficient tool for predicting the distribution of active contaminant in rooms with UFAD.

With the throw comparable to the height of the breathing zone, the inhaled air quality of
seated occupants was similar to the air quality in the lower occupied zone. The ventilation
effectiveness (personal exposure index, pollutant removal efficiency) measured in the
inhaled air was around 2. Loudermilk (2003) stated that mixing (promoted by underfloor
ventilation) must be limited to just below the respiration level of seated occupants in order
to prevent transmission of contaminants between occupants. The present results indicate
that although such stratification decreases the exposures of occupants to one half the
exposure with mixing ventilation, a shorter throw is required in order to reduce the
transmission significantly. Finding the relationship between the inhaled air quality and the
throw height is, however, not easy. The stratification of contaminants as well as mixing of
air in the lower occupied zone (due to UFAD) makes it difficult to assess the ability of the

84
Chapter 5

upward free convection flow around the body to provide clean to the breathing zone (based
on data measured and published so far). Therefore, more research in this area is needed. The
concentration of human-produced contaminants inhaled with different ventilation systems
(without PV) was presented in Figure 5.4.
Concentration (C-CS/CE-CS)

1
UFAD + RMP
0.8
UFAD + VDG
0.6

0.4

0.2

0
0.1 m (65 L/s, 0.3 m (50 L/s) 0.6 m (65 L/s) 1.0 m (80 L/s)
horiz.)
Throw height of UFAD (airflow rate)
Figure 5.30. Concentration of exhaled air from the front manikin inhaled by the back manikin. Back
manikin using PV at 15 L/s while front manikin does not. The data for a throw height of 0.1 m (i.e.
horizontal discharge) and 0.6 m were obtained in stage 2-1. The data for a throw height of 0.3 m and
1.0 m were obtained in stage 2-2. The throw height of 0.6 m, corresponding to the airflow rate of 65
L/s, was not measured but it was estimated as half way between the throw at 50 and 80 L/s.

The inhaled air quality increases substantially when PV is used. The largest improvement is
realized when an occupant inhales air direct from the potential core region of a personalized
air jet. The air quality outside of the core is a product of mixing between clean personalized
air and polluted surrounding air. Figure 5.30 compares the concentration of contaminants
exhaled from the front manikin in the air inhaled by the back manikin at different vertical
throws of UFAD. The back manikin was protected with PV. It is shown that the throw
height of UFAD did not have a substantial impact on the contaminant concentration in the
inhaled air, although the manikin did not inhale direct from the clean core region (if so the
concentration would be zero). Furthermore, the concentrations with PV and UFAD (all
throw heights) were comparable with the concentrations provided by PV and displacement
ventilation (Figure 5.4). The reason is not clear. It is possible that the local concentration
surrounding the air terminals did not change substantially, although the concentration
levels near the floor varied with the throw. The uncertainty of measurements should be
considered as well. In practice, both a high air quality experienced close to the personalized
airflow, and a lower air quality inhaled at the workstation (when PV is not used) will
certainly contribute to the exposures of occupants over a period of time. Hence, a short
throw of UFAD, which decreases the exposures of unprotected occupants (Figure 5.4),
should be prioritized.

In the room with UFAD as well as in the room with displacement ventilation, the use of PV
increased mixing of the contaminants located in its vicinity and thus decreased the inhaled
air quality of other occupants. Cermak and Melikov (2003) observed the same phenomenon
previously; however, only one PV system was studied and the throw height of UFAD was
not considered as a parameter. Figure 5.31 compares the transmission of exhaled air from
the front manikin to the back manikin at different throw heights of UFAD. The front
manikin used PV at 15 L/s, while the back manikin was unprotected. The results for the
RMP showed that the level of transmission did not change with the throw height, unless the
throw was very short. The VDG is shown to cause a lower transmission of contaminants
than the RMP (Cermak and Melikov, 2004). Moreover, the transmission with the VDG

85
Chapter 5

decreases proportionally to the decrease in the throw (discussed below). The direction of
personalized airflow and not its strength is supposed to be the reason, because the airflow
provided upward (VDG) had a greater momentum (and it caused lower transmission) than
the airflow provided downward (RMP). The transmission was somewhat lower with PV and
the horizontal discharge of UFAD than with PV and the displacement ventilation (Figure
5.5). The ability of UFAD to reduce non-uniformities and thus prevent the vertical spread of
contaminants (discussed below) might be the reason.
Concentration (C-CS/CE-CS)

1
UFAD + RMP
0.8
UFAD + VDG
0.6

0.4

0.2

0
0.1 m (65 L/s, 0.3 m (50 L/s) 0.6 m (65 L/s) 1.0 m (80 L/s)
horiz.)
Throw height of UFAD (airflow rate)
Figure 5.31. Concentration of exhaled air from the front manikin inhaled by the back manikin. Front
manikin using PV at 15 L/s, while back manikin does not. The data for a throw height of 0.1 m
(horizontal discharge) and 0.6 m were obtained in stage 2-1. The data for a throw height of 0.3 m and
1.0 m were obtained in stage 2-2.

The extent of mixing and thus the transmission of human-produced contaminants was
expected to decrease with the decrease in the air velocity impinging the source. The ability
of RMP to affect the transmission was examined at a rate of 7 L/s, where its efficiency to
provide clean air to the breathing zone starts to decline (Bolashikov et al., 2003). This
corresponds to a supply air velocity of 0.26 m/s and a Reynolds number of 3200. The
measurements showed, however, that even such a weak airflow was not able to maintain the
transmission low near the reference level without PV (Figure 5.27). Although the non-
uniformities of the human-produced contaminants decreased in the vicinity of the polluting
manikin’s workplace (Figure 5.28), the distribution of the contaminants next to the exposed
manikin did not differ from the case where the supply air velocity was twice as high (at 15
L/s). Contrary to RMP, VDG supplying air at a rate of 7 L/s did not increase significantly
the transmission of contaminants between workplaces as compared to the reference level
without PV. The supply air velocity and the Reynolds number (based on a hydraulic
diameter) were respectively 1.6 m/s and 3900. The analyses showed that the centreline
velocity of the upward airflow from VDG decreases to approx. 0.3 m/s at the ceiling level
(Awbi, 1998), which is still high enough to encourage mixing. Therefore, the ability of PV to
cause mixing of contaminants generated in its vicinity depends more likely on the direction
of PV airflow than on its velocity.

The measurements with UFAD showed that the transmission of exhaled air from the front
manikin to the back manikin and vice versa are comparable, even when the throw of UFAD
is short. The reason can be the ability of UFAD to provide a more uniform contaminant
distribution across the room, which was demonstrated by the reduction of the concentration
peaks in front of the polluting manikin (discussed below). This implies that in rooms with
UFAD, the layout of workplaces will not have an impact on the transmission of
contaminants between occupants (discussed in more detail in Chapter 6).

86
Chapter 5

The ability of PV to affect the distribution of contaminants generated at another workplace is


difficult to generalize. The reason is that the direction, strength, airflow rate and the
construction of supply air terminal were confounded. Table 5.4 summarizes the ability of PV
in combination with UFAD to affect the distribution of contaminants. The physical
properties of personalized air jets are included in the table. It is shown that neither terminal
type affects the distribution of contaminants at a rate of 7 L/s. At higher airflow rates,
however, the ability of the two types of PV to affect room air distribution differs. The RMP
terminal was not able to affect the distribution even at a rate of 30 L/s, while the VDG
caused mixing already at a rate of 15 L/s. The comparison of the physical properties may
suggest that PV does not affect the distribution of the contaminants generated at another
workplace for momentum flux lower than ca. 0.015 N and velocities lower than 1-1.5 m/s,
regardless of the direction of the PV airflow. However, more measurements are needed in
order to identify the threshold values with certainty. The experimental design should
involve several terminals with different dimensions. Each terminal should be tested when
providing air from different directions in order to prevent confounding of terminal
construction and airflow direction (the case in the present study).

Table 5.4. Impact of PV on the distribution of contaminants generated at another workplace.


RMP VDG
Airflow Supply Supply
rate Impact on Momentum velocity Impact on Momentum velocity
(L/s) distribution flux (N) (m/s) distribution flux (N) (m/s)
7 No 0.002 0.26 No* 0.013 1.59
15 No 0.010 0.56 Mixing 0.061 3.41
30 No 0.040 1.12 - - -
* Not presented in detail, observed also with PV and displacement ventilation (Chapter 4)

Room air distribution


In the experiments with displacement ventilation, the use of PV created large non-
uniformities and differences between the distribution of human-bioeffluents and the
distribution of exhaled air in the vicinity of the contaminant source. With UFAD, the
distributions of the two contaminants were similar. The concentration peaks were more than
50% lower in magnitude with the RMP combined with the horizontal discharge of UFAD
(Figure 5.8) than with the RMP combined with the displacement ventilation (Figure 4.12).
With the vertical discharge of UFAD, the peaks vanished completely. The mixing generated
by UFAD, although limited to a relatively short height, reduces thus the contaminant non-
uniformities and the horizontal spread of contaminants in the occupied zone as compared to
displacement ventilation. It is possible that non-uniformities and differences between the
distribution of human bioeffluents and the distribution of exhaled air still exist in the
immediate vicinity of an occupant.

In rooms with PV and UFAD, the distribution of an active contaminant depends on the
interaction between the mixing zone height (throw) and the stratification height. The
stratification height decreases when the airflow rate of UFAD decreases (as with
displacement ventilation). Personalized airflow directed upward entrains surrounding air
and acts like a thermal plume, thus decreasing the stratification height as well. The volume
of air entrained to different PV airflows in a confined office workplace was, however, not
studied. Personalized airflow directed horizontally or downward may not affect the
stratification height; however, it affects the concentration of contaminants in the lower
occupied zone when directed towards the source.

87
Chapter 5

Figure 5.32 illustrates the vertical concentration profiles of active contaminants for three
possible scenarios of a combined performance of PV providing air upward and UFAD
utilizing passive floor diffusers. The concentration profiles are supported by the
measurements (see Figures 5.20-5.23). UFAD alone (without PV) providing a long throw will
reach the height of the stratification and bring contaminants from the upper mixed zone to
the lower zone (Figure 5.32, left). The use of PV will decrease the stratification. If the airflow
rate of UFAD is kept constant when PV is used, the throw and thus the mixing created by
UFAD in the lower zone will not change. The upward airflow from the floor diffusers may
then interact with the decreased stratification, increasing thus the contaminant concentration
in the occupied zone (Figure 5.32, middle). A decrease in the airflow rate of UFAD will
decrease the stratification height even further. At the same time, however, the mixing in the
lower zone may decrease and provide a fairly low concentration level in the lower zone
(Figure 5.32, right). The flow in a human boundary layer may then transport air from the
lower levels upward to the breathing zone, providing a lower exposure of occupants
compared to the former two cases. However, a further decrease in the stratification height
would decrease the inhaled air quality due to the entrainment of polluted surrounding air
by the free convection flow around the body. Because of the complex interaction of
stratification height, throw height and PV airflow, it is premature to elaborate a physical
model that would describe the vertical distribution (and inhaled air concentration) of active
contaminants in rooms with PV and UFAD.

Figure 5.32. Examples of concentration profiles of active contaminants with PV supplying air
upward and UFAD with a vertical discharge. Left: PV is not used - throw is comparable to the height
of the breathing zone. Middle: PV is used - throw height did not change. Right: PV is used – airflow
rate of UFAD and throw height decreased.

The analyses suggest that a design strategy ensuring a low transmission of contaminants
between workplaces should involve a short throw of UFAD and a personalized air terminal
supplying air upward. As long as the throw is short, the control strategy of UFAD is less
important. However, a constant air volume strategy will not decrease the stratification
height (with passive floor diffusers) and will also result in a higher ventilation rate when PV
is used. The airflow rate of UFAD should be chosen to determine a stratification height
according to criteria for displacement ventilation, e.g. comparable to the height of the
breathing zone. The inhaled air quality will not deteriorate due to PV if its airflow rate
(supply velocity) is low, as discussed previously. The worst strategy, on the other hand,
would maintain a long throw, and decrease the airflow rate of UFAD when PV is used. The
throw height of UFAD seems unimportant from the viewpoint of transmission of
contaminants between occupants when a personalized air terminal device generating
horizontal/downward flow is used. However, a shorter throw would provide a lower

88
Chapter 5

exposure of unprotected occupants to other active contaminants present in the room.


Depending on the design of the floor diffusers (discharge pattern), the use of a shorter throw
may be associated with a draught risk and a larger vertical temperature difference between
head and ankles.

Thermal environment
Temperature stratification plays an important role in determining occupants’ thermal
comfort as well as energy performance of the combined systems. With UFAD alone, the air
temperature near the floor increases due to the ability of swirl flows to entrain larger
amounts of surrounding air, as compared to a unidirectional flow of displacement
ventilation. A normalized temperature near the floor measured for the long vertical throw
was 0.7. This corresponds to the temperatures reported in the UFAD design guide by
Bauman and Daly (2003), based on Webster et al. (2002a, b). The normalized temperature
with the vertical discharge providing a short throw and the horizontal discharge was about
0.6. This is much higher than the normalized temperature of 0.4 recorded in the experiments
with displacement ventilation and implies that the risk of thermal discomfort is smaller in
rooms with UFAD. The differences between the temperature stratification with UFAD and
the temperature stratification with displacement ventilation decrease with the distance from
the floor. The shape of the temperature profiles reflects the fact that the majority of the heat
sources are located in the lower occupied zone (e.g. REHVA, 2003). If the heat sources are
located in the upper level of the space, the temperature may not increase sharply near the
floor and the differences between the temperature profiles for the different systems will
probably disappear.

1
Temperature (T-TS/TE-TS)

0.8

0.6

0.4
PV + UFAD
0.2 PV + Displacement
Displacement (Mundt, 1990)
0
0 1 2 3 4 5
2
Room airflow rate (L/s.m )

Figure 5.33. Normalized temperature near the floor (0.1 m) as a function of the room airflow rate per
m2 floor area. Results for displacement ventilation according to Mundt (1990).

The use of PV promoted mixing, disturbed the thermal stratification and thus increased the
temperature in the occupied zone. The temperature in the occupied zone may not, however,
increase if it is already high due to the mixing generated by UFAD. This was the case for the
RMP terminal and UFAD with the vertical discharge. Regardless of the discharge pattern of
UFAD, the temperature increased more with the VDG terminal than with the RMP terminal.
The differences between the terminals, and their impact on the thermal comfort of seated
occupants, are discussed in Chapter 4. Figure 5.33 plots the normalized temperature near the
floor for PV and UFAD (all experiments performed in phase 2-1 and 2-2) versus the overall
room airflow rate. The normalized temperatures for PV and displacement ventilation (all
experiments performed in phase 1) are presented as well. The measurements were taken at a
height of 0.1 m. The curve for displacement ventilation was suggested and verified with a
large number of experimental data by Mundt (1990). It can be seen that the normalized

89
Chapter 5

temperatures with PV and UFAD range between 0.7 and 0.8 over a range of airflow rates.
The values correspond to the highest values of normalized concentration reported by
Bauman and Daly (2003) for underfloor ventilation alone. In their case, however, the vertical
throw of the floor diffusers was generally longer than in the present study (0.6 m to 2 m).
This indicates that the increase in temperature caused by a long throw of UFAD is
comparable to the increase in temperature caused by PV in combination with a shorter
throw. The temperatures with PV and UFAD were higher than the temperatures measured
in rooms with PV and displacement ventilation. The temperature measured with the
displacement ventilation (the lowest diamond in the figure) agrees well with the model of
Mundt (1990).

As discussed in Chapter 4, the model of Mundt (1990) could be used to predict the
normalized temperature near the floor in rooms with PV and displacement ventilation.
However, the model cannot be applied in the case of underfloor ventilation, neither alone
nor in combination with PV. The results presented by Bauman and Daly (2003) for a wide
range of airflow rates suggest that the relationship between the normalized temperature and
the airflow rate does not follow the theoretical relationship as derived by Mundt (1990).
Therefore, more research is needed before any model for practical purposes can be
recommended. In rooms with displacement ventilation, a linear vertical temperature profile
is often assumed as a first approximation. The result for displacement ventilation (discussed
in Chapter 4) showed that the vertical air temperature difference between head and ankles
predicted on the assumption of profile linearity could be very different from the difference
actually measured. An increase in temperature near the floor in rooms with UFAD caused
the linear profile to agree better with the measurement, especially in the case of UFAD
providing a long vertical throw. Therefore, in rooms with UFAD providing a long vertical
throw, the use of a linear profile can be justified. It should be noted, however, that the
temperature profiles measured in the present study were very steep in the lower zone due to
the presence of most heat sources. A good agreement between a linear profile and
measurements can be expected in rooms where the distribution of heat sources is vertically
uniform, regardless of the distribution system.

20
18 B
Mixing
16 Displacement
Draught rating (%)

14 UFAD horizontal (VAV)


12 UFAD vertical (VAV)
10 UFAD vertical short (CAV)
8 UFAD vertical long (CAV)
A
6
4
2
0
no PV with RMP with VDG

Figure 5.34. Draught rating at a height of 0.1m based on the measurement of temperature and
velocity in the room. Categories of thermal environment (A and B) according to CEN (1998) are
indicated.

Draught has been identified as the major local discomfort factor for the occupants in rooms
with displacement ventilation (Pitchurov et al., 2002). Figure 5.34 presents a comparison of
the draught rating with the different ventilation systems tested, with and without PV. The
rating is based on the measurement of mean air temperature, mean air velocity and

90
Chapter 5

turbulence intensity at a height of 0.1 m. The larger value of the rating obtained at positions
A and B (Figure 3.1) is presented. It is shown that in general the use of PV decreased the
draught rating as compared to the reference cases without PV. The highest rating of about
12% was identified for mixing ventilation (due to elevated velocities near the floor, Figure
5.14). Although the rating with displacement ventilation and UFAD providing air
horizontally decreased as compared to mixing ventilation, only UFAD with a vertical air
discharge was able to decrease the draught rating substantially. All values corresponded to
a thermal environment of category A according to CEN Report 1752 (CEN, 1998), as
indicated on the right-hand side of the figure.

10
9 Mixing
Dissatisfied with vert. ∆T (%)

8 C Displacement
7 UFAD horizontal (VAV)
6 UFAD vertical (VAV)
5 UFAD vertical short (CAV)
4 B UFAD vertical long (CAV)
3
2
1 A
0
no PV with RMP with VDG

Figure 5.35. Percentage dissatisfied with a vertical air temperature difference between head and ankles
(1.1 and 0.1 m above the floor). Categories of thermal environment (A, B and C) according to CEN
(1998) are indicated.

The predicted percentage dissatisfied with vertical air temperature difference between head
and ankles is documented in Figure 5.35. The larger value of the vertical temperature
difference obtained at positions A and B (Figure 3.1) was used and converted to the
percentage dissatisfied by means of the relationship presented in CR Report 1752 (CEN,
1998). The lowest percentage dissatisfied was predicted for mixing ventilation and UFAD
providing a long vertical throw. The percentage dissatisfied was low also with all the UFAD
systems. In one case, with UFAD alone (no PV) providing a short vertical throw, the
percentage dissatisfied increased due to the low temperature of supplied air (16.4°C). As
expected, the largest risk of discomfort was identified for displacement ventilation, even
though the supply air temperature was 20°C.

The measurement position is an important factor that determines the prediction of local
thermal discomfort. The measurement positions used in the present study and the thermal
manikins were both located approximately at the same distance from the displacement
diffuser (Figure 3.1). In the case of underfloor ventilation, the distance between the
measurement positions and the floor-based outlets was shorter than the distance between
the manikins and the outlets. Therefore, the results for underfloor ventilation overestimate
the local thermal discomfort of occupants (as compared to displacement ventilation).

The results (Figures 5.34 and 5.35) indicate that UFAD systems with a vertical discharge
pattern have a greater potential to provide a lower risk of local thermal discomfort than
displacement ventilation or UFAD systems with a horizontal discharge pattern, both with
and without PV. On the other hand, displacement ventilation and UFAD with a horizontal
discharge are advantageous in providing a better air quality. Because PV makes it possible
for occupants to increase their inhaled air quality substantially, the ability of a total-volume

91
Chapter 5

ventilation system to provide high air quality may not be important. Therefore, an air
distribution system ensuring a comfortable thermal environment, i.e. UFAD system with a
vertical discharge pattern, should be preferred. Chapter 7 outlines the recommendations for
practical application.

5.5 Conclusions

• The performance of PV in combination with UFAD is affected by the throw height of


UFAD, and the direction and the strength of PV airflow.

• The performance of PV in combination with UFAD providing a short throw is


comparable to the performance of PV in combination with displacement ventilation in
regard to inhaled air quality. However, UFAD is advantageous in decreasing the non-
uniformities of contaminants and the vertical air temperature gradients.

• In the occupied zone of rooms with UFAD providing a short throw, personalized airflow
directed upward will provide a lower concentration of an active contaminant located in
its vicinity than the personalized air directed horizontally/downward. Therefore, the
use of PV with an upward airflow direction may result in a better air quality for other
occupants who are unprotected with PV.

• An increase in the throw height of UFAD decreases the inhaled air quality of occupants,
both with and without PV. The impact of the direction of personalized airflow on the
mixing of contaminants in the occupied zone decreases gradually with the increase in
the throw, and it disappears when the throw is comparable to the height of the breathing
zone.

• The transmission of active contaminants between two workplaces arranged in a row


behind each other is independent of the location of the workplace where the
contaminants are generated. This suggests that the layout and arrangement of an office
may not have a significant impact on the transmission in rooms with UFAD.

92
6
6. General discussion

A series of physical measurements was carried out in order to examine the performance of
personalized ventilation (PV) in combination with different total-volume ventilation
systems. Two types of air terminal device for PV were examined. At present, they represent
the concepts with the highest potential to ensure excellent air quality and preferred thermal
environment for occupants. The major difference between the terminals was the direction of
personalized airflow in relation to the occupant. The total-volume ventilation systems were
mixing ventilation, displacement ventilation and underfloor ventilation. The impact of the
total-volume ventilation principle on the performance of PV, and vice versa, was examined
under well-defined conditions. The methods, the experimental design and the conditions
tested were identical. This is important, because it allows for a credible comparison of the
performance of the combined systems. The criteria for evaluation were the air quality,
assessed on the basis of the distribution of three simulated contaminants (floor covering,
human bioeffluents and exhaled air) and inhaled air temperature, and thermal comfort of
occupants. The results for each combined system are discussed in detail in the
corresponding chapters. The most important findings and their possible implications in
practice are discussed in the following.

Air quality
Ventilation by mixing represents the most common air distribution principle in practice. In
the present study, the mixing ventilation system acted as a reference case for all other
distribution systems. The measurements showed that the mean air velocity in the occupied
zone was high enough to affect buoyancy-generated primary airflows such as a free
convection flow around occupants, thermal plumes above electronic equipment, etc.
Therefore, the distribution of the three contaminants, temperature and velocity were close to
uniform. The use of PV provided a substantial improvement of the inhaled air quality. The
amount of clean personalized air inhaled was influenced by the efficiency of the air terminal
devices and their positioning in respect to the occupant (discussed in Chapter 4). The results
generally agree with the results of the previous studies reported in the literature (Bolashikov
et al., 2003; Melikov et al., 2002; Faulkner et al., 1999). The use of PV in combination with
mixing ventilation did not have an impact on the room ambient environment in any way,
although rather extreme patterns of PV use were tested (PV was used at high airflow rates
by different occupants at a time). Neither did the room airflow affect the performance of PV.
Therefore, as compared to the performance of a mixing ventilation system alone, the
application of PV in rooms with mixing ventilation can only be beneficial.

The present study, the first of its kind, examined the performance of PV in a room with a
controlled non-uniform distribution of contaminants. The non-uniform distribution was
provided with a displacement ventilation system and an underfloor ventilation system.
With the displacement ventilation system alone, the room air distribution followed the
expected pattern. The distribution of the active contaminant (human bioeffluents) was
stratified with a low concentration near the floor and a high concentration near the ceiling.
The interface layer between the two zones was formed at a height of about 1 m. This ensured

93
Chapter 6

a high air quality for seated occupants, whose breathing zone was served with clean air by
means of a free convection flow around the body. No differences were found between the
distribution of bioeffluents and the distribution of exhaled air (exhaled air is heated and
discharged with an initial momentum while bioeffluents are carried by the convection flow
around the body). The distribution of the floor contaminant was generally uniform, except
near the floor. The entrainment of air in the human boundary layer caused the inhaled air
quality, in regard to the floor contaminant, to be comparable to the air quality in the ambient
air at the height of the breathing zone.

The vertical throw of supply airflows determines the distribution of active contaminants in
rooms with underfloor ventilation. The present results showed that the distribution of
contaminants with UFAD providing a short throw was comparable to the distribution of
contaminants with displacement ventilation. This is in agreement with Fisk et al. (1991). An
increase in the throw provided a more intensive mixing, which increased the concentration
of active contaminants in the occupied zone. Analyses showed that a room air distribution
model proposed recently by Yamanaka et al. (2002) agreed rather well with the
measurements obtained in the present study (Appendix F).

The use of PV provided a substantial improvement in the quality of inhaled air. With the
most efficient air terminal device tested, the inhaled air concentration of contaminants
decreased as much as 50 times compared to the concentration of the contaminants in the
exhaust air. However, in a room with displacement or underfloor ventilation, the use of PV
was shown to promote mixing of contaminants located in its vicinity. This is the case of
human bioeffluents, exhaled air (both studied) and possibly also contaminants located on
occupants’ desks. The results indicate that the mixing due to PV may increase the transport
of contaminants between workplaces and decrease the air quality of occupants unprotected
with PV, as compared to the case where PV was not used. This may be of concern in cases
with tobacco smoking, or when occupants are sources of contagious diseases and the control
of their transmission is desirable. However, the air quality of occupants unprotected with
PV was still higher than in a room with mixing ventilation. As discussed in Chapter 4, the
mixing caused by PV could be comparable to mixing caused by an activity of occupants, and
therefore it might not be considered as a drawback of PV systems in real life. Such
comparison, however, remains to be made. It was identified that the inhaled air quality of
unprotected occupants depends on the direction and strength of the personalized airflow
used by other occupants. Personalized air directed upward was shown to ensure a better air
quality than personalized air directed downward or horizontally. The phenomenon was
observed with both displacement and underfloor ventilation (regardless of its throw height).
The differences between the systems, in particular the impact of the throw height of UFAD,
were discussed in detail in Chapter 5.

The use of PV in a room with displacement ventilation or underfloor ventilation may or may
not affect the distribution of active contaminants generated further from the personalized
airflow. Such contaminants can be human bioeffluents and exhaled air generated elsewhere
in a room, contaminants from office equipment or even passive contaminants carried by
means of thermal plumes above various heated objects. The ability of PV not to affect the
distribution of human-produced contaminants may be of interest in e.g. admission areas in
hospitals, where a possible infector (unprotected with PV) frequently comes into contact
with hospital personnel (possibly protected with PV). Analyses discussed in detail in
Chapter 5 suggest that PV may not affect the mixing of contaminants generated at another
workplace at velocities below 1-1.5 m/s. However, because of a limited number of
experiments, the threshold values could not be determined with certainty. If available, such

94
Chapter 6

information would allow designers to restrict the supply airflow rate or optimize the
dimensions of an air terminal device so that the impact of PV on the room air distribution is
negligible.

The discussion presented so far concerned only the distribution of active (heated)
contaminants generated from localized sources. However, many more contaminant sources
are present in indoor environments. In the present study, the distribution of passive
contaminants emitted from a plane source located on the floor was examined as well. This
represents pollution (VOCs) from carpet, PVC or linoleum, which has been associated with a
prevalence of SBS symptoms (Section 1.4). The results showed that the distribution of the
floor contaminant was generally uniform with all the total-volume ventilation systems. With
displacement ventilation and underfloor ventilation, non-uniformities were observed in a
region close to the floor where the supply air current diluted the contaminants to a lower
concentration (discussed in Chapter 4). However, the lower concentrations near the floor did
not have an impact on the inhaled air quality of unprotected occupants. The use of PV
improved the inhaled air quality of occupants substantially, while the room distribution of
the floor contaminants was not affected. This is important, because the use PV will thus
always protect occupants from the contaminants emitted from floors.

Risk of indoor airborne transmission of infectious diseases


The Wells-Riley equation is commonly used to model the risk of indoor airborne
transmission of infectious diseases. Rudnick and Milton (2003) derived an alternative
equation that determines the risk of transmission of infectious diseases using CO2
concentration as a marker for exhaled-breath exposure. This is possible because air exhaled
by people is the vehicle for release of respiratory infectious agents (Section 1.4). Analogous
analyses were performed in the present study (air exhaled from the manikins was marked
with tracer-gases). The concept of the normalized concentration was introduced in the
model in order to quantify the impact of various ventilation systems on the risk of infections.
Instead of the probability of infections, as used in the original Wells-Riley equations, the
reproductive number for an infectious disease in a building environment (RA0) was
calculated. The reproductive number is the number of secondary infections that arise when a
single infectious case is introduced into a population where everyone is susceptible. An
infectious agent can spread in a given population, if RA0 > 1. The greater the value of RA0 the
more likely is the infection to reproduce rapidly. The equations used as well as examples of
application are presented in Appendix G.

The typical values of a normalized concentration of exhaled air contaminants in the air
inhaled by the thermal manikins were used to calculate the basic reproductive number for a
typical infectious disease such as influenza. The normalized concentration of 1, 0.2, 0.05 and
0.7 was taken for respectively mixing ventilation alone, UFAD alone, UFAD plus RMP
protecting the occupant, and UFAD plus RMP not protecting the occupant. A quantum
generation rate of 100 quanta/hour estimated by Rudnick and Milton (2003) was used.
Figure 6.1 presents the basic reproductive number at two ventilation rates of outdoor air, of
which one corresponds to the ventilation rate of outdoor air required by present standards
(CEN, 1998; ASHRAE, 1992). It may be assumed that the recirculation of exhaust air to
supply air is utilized in cases with PV and UFAD, where it would be unrealistic to supply an
airflow rate of as low as 10 L/s per person simultaneously with both systems. The chart is
constructed assuming a presence of 30 persons in the room (chosen arbitrarily) and a time of
exposure of 8 hours.

95
Chapter 6

8
10 L/s per person
Reproductive number, RA0
7
40 L/s per person
6
5
4
3
2
1
0
Mixing UFAD RMP+UFAD RMP+UFAD
(protected) (unprotected)

Figure 6.1. Reproductive number for selected ventilation systems (typical values) at different outdoor
air ventilation rates for influenza (generation rate = 100 quanta/hour).

In the case of mixing ventilation and a supply rate of 10 L/s per person, there is a likelihood
that 7 out of 30 occupants contract influenza after an 8-hour exposure. The number of
possibly infected person decreases to just 2 (1 already and 1 secondary infected) if either the
ventilation rate is increased to 40 L/s per person or an underfloor ventilation system is
employed. It is shown that the use of PV enables the occupants to protect themselves
efficiently from infections. However, the use of PV by other occupants increases the risk of
infections for an occupant unprotected with PV. As shown in Figure 6.1, there is a
probability of as many as 3 new cases of influenza if the infected occupant uses PV while the
other occupants do not use PV for protection, as compared to a still environment with
underfloor ventilation alone. Nevertheless, the risk is always lower than in a mixing
ventilated room, and perhaps comparable to the case when an activity of occupants causes
mixing of indoor air.

In the model presented it has been assumed that the loss of viability, filtration, settling and
other mechanisms are small compared with removal by ventilation (discussed by Rudnick
and Milton, 2003). An additional assumption is that the transmission of infectious agents
due to the use of PV is independent of the room size and the space distribution of occupants.
The CFD simulation by Murakami et al. (1992) indicated that in large open-plan offices with
underfloor ventilation, a horizontal spread of contaminants could be reduced when the
supply and exhaust airflow rates of underfloor ventilation are balanced locally in the small
place allocated to each supply opening. The resulting floor-to-ceiling pattern could remove
heat (studied by Murakami et al., 1992) and possibly contaminants near the workstation
where they are generated. This would reduce the horizontal transport of contaminants in
rooms.

The infectiousness of a disease (i.e. estimation of the generation rate of infectious doses) is
crucial for the risk of transmission analyses. With a less infectious agent such as rhinovirus,
the infection would not spread in any of the cases presented in Figure 6.1. This implies that
the ability of PV to increase the transmission of contaminants between occupants does not
pose a problem. On the other hand, it is very difficult to prevent the spread of a highly
infectious disease such as measles. It was calculated for the given conditions (30 occupants,
10 L/s per person) that the spread of measles would cease at a normalized concentration of
0.02 (ventilation effectiveness of 50). This is comparable to the top performance of the RMP
terminals tested. This suggests that PV could be effective in protecting occupants even in
environments with a high risk of infections, and it could become an alternative or
supplement to traditional methods of occupant protection such as high-efficiency filtration

96
Chapter 6

or ultraviolet disinfection of air. In the present study, the highest performance of RMP was
measured in the rooms with displacement and underfloor ventilation. The performance of
RMP in combination with mixing ventilation was somewhat lower. However, as
documented by Bolashikov et al. (2002), the performance of RMP in combination with
mixing ventilation can be extremely high as well (normalized concentration in inhaled air of
0.1, i.e. almost 100% clean air). Therefore, the ability of PV to protect occupants from
infectious diseases can be considered independent of a total-volume ventilation principle.

In the present study, the airborne transmission of respiratory infections associated with
exhaled breath was simulated by means of tracer-gases. In real life, not all droplets
generated in the respiratory tract by talking, sneezing or coughing are airborne immediately
after expulsion. As reviewed in Section 1.4, the size distribution of exhaled droplets may
vary between 0.01 µm to 100 µm or more. After expulsion, the larger droplets either settle
out or evaporate to droplet nuclei (droplets contain impurities) that approach the size of the
individual agent (viruses range between 0.003 and 0.06 µm). The distribution of particle size
is difficult to measure because the droplet size changes in space and time due to evaporation
and coagulation. It is reasonable to assume that most exhaled air droplets smaller than ca. 10
µm (larger droplets are likely to settle out after expulsion) will become airborne and can be
transported by air routes over long distances (see Section 1.4). Therefore, the use of tracer-
gases for the simulation of airborne infection transmission is justified. However, tracer-gases
may not simulate properly the transport of larger droplets, which do not follow air currents
due to momentum (after expulsion) and/or forces of gravity. In real life, both large droplets
and smaller droplet nuclei contribute to the exposure of occupants to infectious agents.

Thermal comfort
The thermal environments evaluated in the present study were not under thermostatic
control. The supply air was conditioned in order to provide a constant cooling capacity
sufficient for the removal of internal heat gains of 580 W (22.5 W/m2). The room air
temperature was 26°C and uniform with the mixing ventilation. With the displacement
ventilation and the underfloor ventilation the temperature distributions were stratified.
Although the inhaled air quality was similar with the two principles (UFAD providing a
short throw), the normalized (dimensionless) temperatures near the floor increased from
about 0.4 with displacement ventilation to 0.6-0.7 with underfloor ventilation. The
normalized temperatures with underfloor ventilation agree well with the previous studies
(Bauman and Daly, 2003; Webster et al., 2002a, b). As discussed in Chapter 5, the increase in
the temperature decreased the risk of draught and a high vertical air temperature difference.

A recent study dealing with a human response to PV systems (Kaczmarczyk, 2003;


Kaczmarczyk et al., accepted) suggested that thermal comfort is an important parameter for
people’s preferences. They showed that occupants take into account both air quality and
thermal comfort when adjusting PV upon arrival at a workplace. With the time elapse, as
occupants became adapted to air quality, thermal comfort started to play a more important
role in determining the individual adjustments of PV. Two thermal manikins were used in
the present study to identify the impact of PV on thermal comfort of seated occupants.
Because the performance of PV is influenced to a large extent by the direction and airflow
rate of personalized air, which are both under occupants’ control, only a general overview of
the topic was given in Chapter 4. The cooling effect of PV was shown to be independent of
the room air distribution principle, i.e. PV affected the thermal comfort the same way when
combined with e.g. mixing ventilation or displacement ventilation. The equivalent
temperature was lower with PV and displacement ventilation than with PV and mixing

97
Chapter 6

ventilation due to the thermal stratification (and identical conditions of supply air). In rooms
with a thermostatic control of the ambient environment the differences would, however,
vanish.

PV did not have an impact on the air temperature field in the room with mixing ventilation.
Although the manikin-based equivalent temperature decreased when PV was used, the
distribution of air temperature measured near the workplaces did not change. The mixing
ensured that the heat removal capacity of both personalized and total-volume ventilation air
spread uniformly. The situation was different in a room with displacement and underfloor
ventilation, where the heat is removed by means of thermal plumes. As documented by the
measurement of velocity, personalized airflow did not affect the air pattern in the lower
occupied zone, and hence it did not contribute to the removal of heat from the sources
located near the floor. This led to an increase in the temperature in the present study,
because either the airflow rate of total-volume ventilation decreased (VAV control) or the
supply air temperature of total-volume ventilation increased (CAV control) when PV was
used. The ability of PV to promote mixing also contributed to the increase in temperature in
the occupied zone.

In practice, a constant air temperature is usually maintained in the occupied zone. Because
the temperature is likely to increase in rooms with thermal stratification when PV is used,
more energy will be needed if a constant room air temperature should be ensured. A higher
airflow rate or a lower supply air temperature will be required if a total-volume ventilation
system is controlled according to a VAV strategy or a CAV strategy, respectively. The room
air temperature and thus the energy consumption may not, however, increase substantially.
The measurements showed that the largest increase in the air temperature was about 1°C at
a height of 0.6 m, and around 0.5°C at another heights, when all PV units were used at a
high airflow rate. When the personalized airflow rate decreased or only one unit was used,
the increase in the temperature was small, comparable to the uncertainty of measurement.
The increase in the temperature does not pose a problem for the thermal comfort of
occupants either, because the cooling effect of PV is large compared to the change of the
ambient environment. Because the increase in the temperature was small, a chain
adjustment of PV by individuals (in order to compensate for continuous environmental
changes) is not likely to occur. To the contrary, an increase in the temperature near the floor
is advantageous in decreasing the risk of draught and vertical temperature difference.

Office arrangement
The layout and furnishing of an office may have a great influence on the airflow pattern and
thus the distribution of contaminants. Nonetheless, considering the large number of possible
arrangements, evaluation of the impact of office arrangement is problematic and difficult to
generalize. In the present study, the workplaces were arranged behind each other with the
occupants facing in the same direction. An active contaminant was generated in the front
workplace in most experiments, while the exposure of an occupant seated at the back
workplace was evaluated. It was assumed that the use of PV would, due to its airflow
direction, carry contaminants from the front workplace to the back workplace, and thus
provide unfavourable conditions (high ambient concentration) for the PV. Several
experiments with underfloor ventilation (stage 2-2) showed that the transmission of active
contaminants between the two workplaces was independent of the location of the workplace
where the contaminants were generated. This suggested that the layout and arrangement of
an office might not have a large impact on the transmission in rooms with UFAD. In rooms
with displacement ventilation, contaminants may spread horizontally over relatively long
distances, and the impact of the office arrangement should therefore be considered. The

98
Chapter 6

workplaces were not partitioned in the present study. The use of partitions is expected to be
beneficial, because they may restrict the horizontal movement of air and thus decrease the
transport of contaminants between workplaces. It is also possible that partitions will restrict
the movement of clean personalized air from workplaces. This may improve the exposures
of occupants when they do not inhale clean air direct from PV airflow.

Air quality in terms of different quantities


Different quantities have been used and reported in the literature to evaluate the exposure of
occupants to contaminants. In the present study the normalized concentration of the
contaminant, defined as the concentration of a contaminant at a point divided by the
concentration of a contaminant in the exhaust (Section 3.7.1), was used for its simplicity. The
calculation of two other indices used frequently – ventilation effectiveness and personal
exposure effectiveness – is straightforward using the normalized concentration.

The reciprocal value of the normalized concentration has been referred to as the ventilation
effectiveness, εV, (CEN, 1998) or the pollutant removal efficiency, PRE (e.g. Faulkner et al.,
1999). With the aim of differentiating between the mean ventilation effectiveness in the
occupied zone and the ventilation effectiveness at different locations, Brohus and Nielsen
(1996) introduced indexes named the ventilation effectiveness in the occupied zone, the local
ventilation effectiveness (for a point in the room) and the personal exposure index (for
inhaled air). Despite the different definitions, the value of a ventilation effectiveness
indicates how many times the concentration at a point or an area of interest is lower that the
concentration in the exhaust. Moreover, the value of ventilation effectiveness shows how
many times more clean air would be needed for a mixing ventilation system alone to achieve
the same concentration of contaminants as the ventilation system in question. For example,
in order to achieve the same level of air quality as RMP (ventilation effectiveness of 50), a
mixing ventilation system alone would have to supply a 50 times higher airflow rate than
RMP in combination with mixing ventilation. However, such airflow rates (provided with
mixing ventilation) would obviously cause draught and increase the consumption of energy.

The problem of ventilation effectiveness is the fact that it converges towards infinity when
the concentration at the point approaches the supply air concentration. The index is thus not
suitable for PV applications, where high amounts of clean air (i.e. very low concentrations)
are expected in occupant’s inhalation, as demonstrated by Melikov et al. (2002). On the other
hand, the ventilation effectiveness is equivalent to the dilution factor as used by Haghighat
et al. (2001). They showed that the acceptability of air is proportional to the logarithm of the
dilution factor (ventilation effectiveness). If extended with the impact of enthalpy on the
acceptability of air (Fang et al., 1998a, b), as suggested by Melikov et al. (submitted), the
transformation could be used for the assessment of perceived air quality with PV. The
logarithm relation between the acceptability and the ventilation effectiveness suggests that
an order of magnitude increase of the ventilation effectiveness is needed in order to yield an
improvement of the perceived air quality. The largest values of ventilation effectiveness
identified in the present study were 50 (normalized concentration of 0.02). This indicates
that a substantial improvement of the perceived air quality, and hence a decrease in the
prevalence of SBS symptoms, may be possible with PV. This was already indicated in the
example of selected SBS symptoms in the experiments with human subject (Kaczmarczyk et
al., 2002a, accepted; Kaczmarczyk, 2003).

The concept of the personal exposure effectiveness index, proposed by Melikov et al. (2002),
was described in detail in Section 3.7.1. The index expresses the portion of clean air from a
ventilation system in the inhaled air, as compared to a reference case. The index makes it

99
Chapter 6

possible to evaluate the contribution of different combined systems to the inhaled air
quality. This is demonstrated in Table 6.1. The normalized concentrations, which were used
for the calculation of the personal exposure effectiveness by means of Equation 3.5, are
presented in the table as well (taken from Figure 5.4). With mixing ventilation alone the
inhaled air consisted entirely of room air (i.e. personal exposure effectiveness of 0). The
inhaled air consisted of 85% clean personalized air when PV was used in combination with
mixing ventilation. The use of PV in combination with UFAD increases the portion of clean
air inhaled from both PV and UFAD to 98%, which is close to the aim of 100% clean air in
inhaled air.

Table 6.1. Personal exposure effectiveness with different ventilation systems.


Ventilation principle Normalized conc. (-) Personal Exp. Effect. (%)
Mixing 0.94 0
PV + mixing 0.14 86
PV + UFAD 0.02 98

In outdoor environmental health literature, the concept of intake fraction (iF) has been used
for comparative risk assessment. The index expresses the fraction of a pollutant release that
is inhaled by a receptor (Bennett et al., 2002; Marshall et al., 2002). In indoor environments,
iF depends on the ventilation rate, the effectiveness of ventilation (normalized
concentration), the breathing rate of occupants, the number of occupants and the duration of
exposure (Appendix H). Figure 6.2 presents the individual intake fraction (single occupant)
with different ventilation systems. The data are based on the results obtained in the present
study for a human-produced contaminant (bioeffluents). The normalized concentrations of
1, 0.2, 0.05 and 0.7 were used respectively for the four cases presented. In the first approach
(relative comparison of systems) it was assumed that an occupant spends 100% of time
indoors. The breathing rate of an occupant was assumed to be 6 L/min.

12000
10 L/s per person
10000 40 L/s per person
iF (per million)

8000

6000

4000

2000

0
Mixing UFAD RMP+UFAD RMP+UFAD
(protected) (unprotected)

Figure 6.2. Intake fraction for indoor environments with different ventilation systems.

Figure 6.2 shows that the intake fraction may range between 125 and 10000 per million
depending on the ventilation system and the ventilation rate of outdoor air. The largest
value corresponds to a room with mixing ventilation providing a ventilation rate of 10 L/s
per person, which is required by standards nowadays (CEN, 1998; ASHRAE, 1989).
Providing that occupants are exposed to an office environment 23% of the time annually (8
hours x 5 days x 50 weeks), the annual average intake fraction with mixing ventilation
decreases to 2300 per million. This is, however, about one order of magnitude the exposure
to outdoor sources in urban areas, which ranges between 1-100 per million (for summary of
different studies see Marshall et al., 2002). This underlines the importance of occupants’

100
Chapter 6

exposures to indoor sources, which may be much larger than the exposures to outdoor
source (e.g. tobacco smoking vs. emission of motor vehicles). Furthermore, the wide spread
of the values of the intake fraction for indoor environments documents the importance of
outdoor ventilation rates as well as the efficiency of ventilation systems. The impact of
various contaminants is included in the normalized concentration, which evaluates the non-
uniformity of their distribution. The analyses indicate that the exposure levels provided with
PV can be comparable to the exposure levels outdoors.

The potential adverse effect from a pollutant release depends not only on the intake of
individuals, as expressed by the intake fraction, but also on the likelihood of adverse effects
with increasing intake. The likelihood of adverse effects, such as a spread of contagious
diseases, was demonstrated and discussed previously. As described in Appendix G, the
concept of intake fraction was actually used in the model by Rudnick and Milton (2003),
referred to as the “volume fraction of inhaled air that is exhaled breath”.

Murakami et al. (2000) proposed two indices that express the Contribution Ratio of a
Pollution (CRP) source. The CRP 1 indicates the percentage of a pollutant (e.g. VOC) inhaled
after being released from the source. The concepts of the CRP 1 and the intake fraction are
thus identical. Murakami et al. (2000) documented the concept of the CRP in a case study.
Computer simulation was performed in a room in order to determine the distribution of
contaminants emitted from the walls, floor and ceiling. Exposure of an occupant standing in
the centre of the room was determined. The room was ventilated by a displacement
principle. The CRP 1 was about 2.6-2.8% (or 26000-28000 per million) for the floor area, while
for the ceiling it was 0-0.09% (0-900 per million). Because the CRP 1 depends on the
ventilation rate, the breathing rate and the efficiency of air distribution, it is difficult to make
a comparison with the present results without analyzing the airflow patterns in detail. The
CRP 2 indicates the portion of pollution from a source with the total amount of pollution
inhaled taken as 100%. In the case study by Murakami et al. (2000), 65% of all contaminants
inhaled were the contaminants emitted from the floor, and 10% were the contaminants
emitted from the ceiling. This implies that the pollution emitted from the floor is more
important than the pollution emitted from the ceiling. Such analyses may be an effective tool
for the evaluation of personal exposures; however, measurements with numerous sources or
complex CFD analyses are required.

Applicability
The ratio of personalized airflow rate to total-volume ventilation airflow rate may be
important for the ability of PV to affect the distribution of contaminants and the thermal
stratification. It might be expected that the impact of PV on the room air distribution will
increase when the ratio of the airflow rates increases, i.e. the personalized airflow rate is
relatively large as compared to the total-volume ventilation airflow rate. However, no
correlation between the ratio of the airflow rates and e.g. the transport of contaminants
between occupants was found. The direction and strength of personalized airflow was
found to be more important. For example, a personalized airflow from VDG terminals was
able to affect the distribution of contaminants generated at another workplace when the
ratio was small (e.g. 15/80 = 0.19), while airflows from RMP terminals did not affect the
distribution when the ratio was high (e.g. 30/50 = 0.6). The limit case is a room with PV
alone (total-volume ventilation may be natural). The impact of one type of PV (PEM, see
Section 1.3) on the distribution of temperature and velocity, operated in a room without any
other ventilation system except PV, was reported in Bauman et al. (1993); however, no
generalization was attempted.

101
Chapter 6

Studies by Kaczmarczyk et al. (2002b, accepted) showed that occupants used efficiently the
delegated control of the airflow rate and the positioning of the air terminal devices. Some
occupants may choose not to use the PV system, while others would exploit the system to its
maximum capacity. Such extreme conditions in terms of a PV supply rate have been
considered in this study. The impact of factors such as the distribution of heat sources, total
heat gain, air change rate, etc. has not been studied. It is assumed that the present findings
will apply for various situations in typical offices, where total-volume ventilation provides a
recognizable and well-defined airflow pattern in the room.

102
7
7. Recommendations

The ability of occupants to generate and control their preferred environment is an important
feature of PV systems. Because occupants may choose not to use PV, or they may use PV in a
way that does not provide a substantially better air quality but preferred thermal comfort,
the quality of the background environment must be considered. The present results
identified that each combination of PV and total-volume ventilation tested can be applied in
practice. This is an important result, because many other factors, such as an architectural
concept, energy saving requirements or preferences of the building owner, may affect the
selection of a total-volume ventilation system in real life.

The use of PV in rooms with mixing ventilation may only be beneficial for both air quality
and thermal comfort of occupants. The inhaled air quality improves in rooms with
displacement ventilation or underfloor ventilation. Although the use of PV promotes mixing
of contaminants generated in its vicinity, the inhaled air quality in rooms with displacement
ventilation and underfloor ventilation can be better (depending on the extent of mixing)
than in a room with mixing ventilation. The inhaled air quality in regard to a plane
contaminant located on the floor does not differ with the different ventilation systems. The
inhaled air quality in regard to active contaminants is higher with displacement ventilation
and UFAD providing a horizontal discharge pattern than with UFAD providing a long
throw. The ranking of the systems in terms of the risk of local thermal comfort is, however,
opposite. Both displacement ventilation and UFAD with a horizontal discharge provide low
temperatures and high velocities near the floor. An increase in the throw height of UFAD
increases temperatures and decreases the velocities, which decreases the risk of draught. The
present study identified that UFAD with a vertical throw of about 0.3 m provides as high air
quality as displacement ventilation, but provides almost as a low risk of draught as UFAD
with a long throw (Figures 5.34). The use of UFAD providing a short throw is thus
recommended. The control of supply air conditions is less important as long as the throw is
short. A constant air volume (CAV) strategy may be preferable in combination with passive
floor diffusers, because variations in airflow rate (VAV strategy) may increase the throw
height and thus decrease the air quality. Moreover, the use of PV will increase the total
amount of ventilation air provided when a CAV strategy is utilized.

The ventilation systems tested in the present study provided 100% outdoor air. The primary
reason was to increase the sensitivity of tracer-gas measurements. In practice, recirculation
of exhaust air to supply air is often used, because it is desirable to keep the quantity of
outdoor air to a minimum for economic reasons and to conserve energy. A certain economic
benefit could be achieved if only a PV system provided outdoor air (in accordance with the
present standards), while another (total-volume air distribution) system would provide
conditioning of room air. If the recirculation were utilized, the shape of normalized
concentration profiles would not change from the situation when 100% outdoor air is
supplied; however, the absolute concentration of contaminants in a room would increase
proportionally to the dilution of contaminants. A cross-sectional questionnaire and field
study (Sundell et al., 1994) showed no increased risk of SBS associated with recirculation of

103
Chapter 7

air; however, they found an elevated prevalence of general SBS symptoms with outdoor air
supply rates lower than 13.6 L/s per person. Another study (Milton et al., 2000) found a
consistent association of increased sick leave with lower levels of outdoor air supply and
indoor air quality complaints. Therefore, the use of recirculation may be acceptable (ambient
air quality), so long as the outdoor air ventilation rate is high. The use of PV will provide
benefits in terms of the improvement of air quality and thermal comfort of seated occupants.

Air terminal devices representing the most promising solutions for providing excellent air
quality and preferred thermal comfort for occupants were examined in the present study.
However, new air terminal devices are being constantly developed, tested and optimized. It
was shown that the impact of PV on the air quality and thermal comfort of occupants is in
general very localized. As indicated by this and other studies (e.g. Faulkner et al., 1993), PV
discharging air at high velocities causes mixing. This affects the upward air movement in
rooms with displacement ventilation and underfloor ventilation and increases temperature
in the lower occupied zone. Providing personalized air at lower velocities, e.g. less than 1-1.5
m/s, is thus desirable and recommended. Such velocities still provide occupants with the
ability to adjust their thermal comfort according to their preferences.

The airflow direction is important when personalized air is directed towards a contaminant
source. If so, the upward direction is preferable to the horizontal/downward direction,
regardless of the personalized airflow rate. The preferable air terminal device should be
highly efficient while not increasing the transmission of contaminants between workplaces.
However, none of the air terminal devices tested in the present study had such qualities. The
RMP terminal was extremely efficient (providing almost 100% clean air in inhalation), but it
caused an increase in the transmission of human-produced contaminants between
workplaces. The VDG terminal, on the other hand, did not increase the transmission, but its
efficiency was lower. Therefore, the development of a new air terminal device is desirable.
Because indoor air is polluted from a large number of sources, and not only by human
bioeffluents and exhaled air, the efficiency of air terminal devices should be preferred to the
their ability to cause mixing. At present, the use of the RMP terminal devices can be
recommended.

Table 7.1. Comparison of air quality and thermal comfort performance of the ventilation systems
tested. Scale: 1 to 5 (1-worst, 5-best).
Total-volume ventilation No PV VDG / RMP
Air quality Thermal Air quality Thermal
comfort Protected Unprotected comfort
occupant occupant
Mixing 1 3 3/4 1/1 4/5
Displacement 3 1 4/5 3/2 2/3
Underfloor horizontal 3 1-2 4/5 4/3 3/4
Underfloor vertical short 3 2-3 4/5 3/2 4/5
Underfloor vertical long 2 3 4/5 2/1 4/5

Table 7.1 provides a comparison of total-volume ventilation systems alone and in


combination with PV tested in the present study. Air quality and thermal comfort of
occupants were the criteria. With PV, air quality of occupants protected and unprotected
with PV is considered. Air quality of occupants unprotected with PV is associated with the
ability of PV used by other occupants in the room to cause transmission of contaminants
between workplaces.

104
8
8. Conclusions

The performance of personalized ventilation in combination with total-volume ventilation


comprising mixing, displacement and underfloor air distribution was studied in regard to
the air quality and thermal comfort of occupants. The main findings are as follows:

• Personalized ventilation is able to protect occupants from pollution and thus increase the
quality of inhaled air in comparison with total-volume ventilation principles. The
positioning of an air terminal device in respect to the occupant and its design influences
the quality of inhaled air to a large extent.

• The use of PV promotes mixing of contaminants located in its vicinity. In rooms with
displacement or underfloor air distribution, this may cause an increase in the
transmission of contaminants between workplaces and thus decrease the air quality of
occupants as compared to the case without PV. Personalized air directed upward
ensures a better air quality than personalized air directed horizontally/downward.

• In rooms with displacement or underfloor air distribution, the use of PV does not affect
the distribution of active contaminants generated in another workplace when the supply
airflow rate is low. At a high rate the type of air terminal device and its airflow direction
may be important.

• The cooling of occupants provided with PV is rather independent of the room air
distribution generated by a total-volume ventilation system. The impact of PV on the
distribution of temperature and velocity in the room is very localized.

• PV does not contribute to the removal of heat from the lower occupied zone in rooms
with displacement or underfloor air distribution. This may increase the temperature in
the occupied zone and consequently the consumption of energy, if the temperature
should be maintained constant.

• The analyses indicate that PV in combination with any total-volume ventilation principle
could be efficient in protecting occupants even from highly infectious diseases, and
therefore become an alternative or supplement to traditional methods of occupant
protection.

The following recommendations can be outlined:

• A combination of PV and underfloor air distribution with a vertical discharge direction


providing a short throw, controlled according to a CAV strategy, is recommended. If the
control of an airborne transmission of contaminants between occupants is desirable, air
terminal devices supplying air upward are preferable.

• Development of air terminal devices with a high efficiency and a low ability to promote
mixing of contaminants located in its vicinity is recommended.

105
References

Aksenov, A.A., Gudzovski, A.V., Shilkrot, E.O. and Zhivov, A.M. (1998) “Thermal plumes
above heat sources in rooms with a temperature stratification”, Proc. of Roomvent 1998,
14-17 June, Stockholm, Sweden, pp. 437-444.
Arens, EA., Bauman, FS., Johnston, LP. and Zhang, H. (1991) “Testing of localized
ventilation systems in a new controlled environmental chamber”, Indoor Air 3, pp. 263-
281.
ASHRAE (1989) “ASHRAE Standard 62: Ventilation for acceptable indoor air quality”,
American Society of Heating, Refrigerating and Air-Conditioning Engineers, Atlanta.
ASHRAE (2000) “Handbook: HVAC systems and equipment”, American Society of Heating,
Refrigerating and Air-Conditioning Engineers, Atlanta, ISBN 1-883413-81-8.
Awbi H. (1998) “Ventilation of buildings”, E & FN Spon, ISBN 0-419-21080-6.
Bakó-Biró, Z., Wargocki, P., Weschler, C.J. and Fanger P.O. (2004) “Effects of pollution from
personal computers on perceived air quality, SBS symptoms and productivity in offices”,
Indoor Air 14(3), pp. 178-187.
Bauman, F.S., Johnston, L.P., Zhang, H. and Arens, E.A. (1991) “Performance testing of a
floor-based, occupant controlled office ventilation system”, ASHRAE Transactions, Vol.
97(1), pp. 553-565.
Bauman, F.S., Zhang, H., Arens, E.A. and Benton, C.C. (1993) “Localized comfort control
with a desktop task conditioning system: laboratory and field measurements”, ASHRAE
Transactions, Vol. 99(2), pp 733-749.
Bauman, F.S., Carter, T.G., Baughman A.V. and Arens, E.A. (1998) “Field study of the impact
of a desktop Task/Ambient Conditioning System in office buildings”, ASHRAE
Transactions, Vol. 104(1), pp. 125-142.
Bauman, F. and Webster, T. (2001) “Outlook for underfloor air distribution”, ASHRAE
Journal, Vol. 43, No. 6 (June), pp. 18-27.
Bauman, F.S. and Daly, A. (2003) “Underfloor air distribution design guide”, ASHRAE, ISBN
1-931862-21-4.
Bennett, D.H., Margni, M.D., McKone, T.E. and Jolliet, O. (2002) “Intake fraction for
multimedia pollutants: a tool for life cycle analysis and comparative risk assessment”,
Risk Analysis, Vol. 22, No. 5, pp. 905-918.
Bjørn, E. (2002) “Dispersal of exhaled air in stratified surroundings – CFD studies”, Proc. of
Roomvent 2002, 8-11 September, Copenhagen, Denmark, pp. 285-288.
Bjørn, E. and Nielsen, P.V. (1996) “Exposure due to interacting air flows between two
persons”, Proc. of Roomvent 1998, 17-19 July, Yokohama, Japan, pp. 107-114.
Bjørn, E., Mattsson, M., Sandberg, M. and Nielsen, P.V. (1997) “Displacement ventilation –
effects of movement and exhalation”, Proc. of Healthy Buildings 1997, 27 September-3
October, Washington DC, USA, pp. 163-168.
Bjørn, E. and Nielsen, P.V. (1998) “CFD simulations of contaminant transport between two
breathing persons”, Proc. of Roomvent 1998, 14-17 June, Stockholm, Sweden, pp. 133-140.
Bluyssen, P.M., de Oliveira Fernandes, E., Groes, L., Clausen, G., Fanger, P.O., Valbjørn, O.,
Bernhard, C.A., Roulet, C.A. (1996) “European indoor air quality audit project in 56
office buildings”, Indoor Air 6, pp. 221-238.
Bolashikov Z., Nikolaev L., Melikov A., Kaczmarczyk K. and Fanger P.O. (2003) “New air
terminal devices with high efficiency for personalized ventilation application”, Proc. of
Healthy Buildings 2003, 4-11 December, Singapore, pp. 850-855.
Brohus, H. (1997) “Personal exposure to contaminant sources in ventilated rooms”, Ph.D.
thesis, Aalborg University, Denmark, ISBN/ISSN: 0902-7953 (R9741).

107
References

Brohus, H. and Nielsen, P.V. (1994) “Contaminant distribution around persons in rooms
ventilated by displacement ventilation”, Proc. of Roomvent 1994, 15-17 June, Krakow,
Poland, pp. 293-312.
Brohus, H. and Nielsen, P.V. (1995) “Personal exposure to contaminant sources in a uniform
velocity field”, Proc. of Healthy Buildings 1995, 10-15 September, Milano, Italy, pp. 1555-
1560.
Brohus, H. and Nielsen, P. (1996) “Personal exposure in displacement ventilated rooms”,
Indoor Air 6, pp. 157-167.
Brosseau, L.M., Vesley, D., Kuehn, T.H., Goyal, S.M., Chen, S. and Gabel., C.L. (1994)
“Identification and control of viral aerosols in indoor environments”, ASHRAE
Transactions, Vol. 100(2), pp. 368-379.
CEN (1998) “CR 1752: Ventilation for buildings – design criteria for the indoor
environment”, European Committee for Standardization, Brussels.
Cermak, R., Holsøe, J., Meyer, K.E. and Melikov, A.K. (2002) ”PIV measurements at the
breathing zone with personalized ventilation”, Proc. of Roomvent 2002, 8-11 September,
Copenhagen, Denmark, pp. 349-353.
Cermak, R. and Melikov A.K. (2003) “Transmission of exhaled air between occupants in
rooms with personalized and underfloor ventilation”, Proc. of Healthy Buildings 2003, 4-11
December, Singapore, pp. 486-491.
Cermak R. and Melikov A.K. (2004) “Transmission of exhaled air between occupants in
rooms with personalized and underfloor ventilation”, Proc. of Roomvent 2004, 5–8
September, Coimbra, Portugal.
Cermak R., Melikov A.K., Forejt L. and Kovar O. (2004) “Distribution of contaminants in the
occupied zone of a room with personalized and displacement ventilation”, Proc. of
Roomvent 2004, 5–8 September, Coimbra, Portugal.
Chiang, H., Su, C.C., Pan, C.S. and Tsau, F.H. (2002) “Study of an innovative partition-type
personal modulation air-conditioning system”, Proc. of Indoor Air 2002, 30 June-5 July,
Monterey, California, pp. 289-294.
Cho, S.H., Kim, W.T. and Zaheer-uddin, M. (2001) “Thermal characteristics of a personal
environment module task air conditioning system: an experimental study”, Energy
Conversion and Management 42, pp. 1023-1031.
Duguid, J.P. (1945) “The size and the duration of air-carriage of respiratory droplets and
droplet-nuclei”, J. Hygiene 54, pp. 471-479.
Etheridge, D. and Sandberg, M. (1996) “Building ventilation: theory and measurement”, John
Wiley & Sons, ISBN 0-471-96087-X.
Fairchild, C.I. and Stamper, J.F. (1987) “Particle concentration in exhaled breath”, Am. Ind.
Hyg. Assoc. J. 48, pp. 748-949.
Fang, L., Clausen, G., Fanger, P.O. (1998a) “Impact of temperature and humidity on the
perception of indoor air quality”, Indoor Air 8, pp. 80-90.
Fang L, Clausen G, and Fanger P.O. (1998b) “Impact of temperature and humidity on the
perception of indoor air quality during immediate and longer whole-body exposures”,
Indoor Air 8, pp. 276-284.
Fanger, P.O. (1998) “Discomfort caused by odorants and irritants in the air”, Indoor Air
Suppl. 4, pp. 81-86.
Fanger, P.O. (2001) “Human requirements in future air-conditioned environments”, Int.
Journal of Refrigeration 24, pp. 148-153.
Fanger, P.O., Melikov A.K., Hanzawa, H. and Ring, J. (1988) “Air turbulence and sensation
of draught”, Energy and Buildings 12, pp. 21-39.
Faulkner, D., Fisk, W.J. and Sullivan, D.P. (1993) “Indoor airflow and pollutant removal in a
room with desktop ventilation”, ASHRAE Transactions, Vol. 99(2), pp. 750-758.

108
References

Faulkner, D., Fisk, W.J., Sullivan, D.P. and Wyon, D.P. (1999) “Ventilation efficiencies of
desk-mounted task/ambient conditioning systems”, Indoor Air 9, pp. 273-281.
Faulkner, D., Fisk, W.J., Sullivan, D.P. and Lee, S.M. (2002) “Ventilation efficiencies of a
desk-edge-mounted task ventilation system”, Proc. of Indoor Air 2002, 30 June-5 July,
Monterey, California, pp. 1060-1065.
Fisk, W.J., Faulkner, D., Pih, D., McNeel, P.J., Bauman, F.S. and Arens, E.A. (1991) “Indoor
air flow and pollutant removal in a room with task ventilation”, Indoor Air 3, pp. 247-262.
Fisk, W.J., Faulkner, D., Sullivan, D. and Bauman, F. (1997) “Air change effectiveness and
pollutant removal efficiency during adverse mixing conditions”, Indoor Air 1, pp. 55-63.
Forejt, L., Melikov, A.K., Cermak, R. and Kovar, O. (2004) “Thermal comfort of seated
occupants in rooms with personalized ventilation combined with mixing or
displacement ventilation”, Proc. of Roomvent 2004, 5–8 September, Coimbra, Portugal.
Haghighat, F., Sakr, W., Gunnarsen, L. and Von Grunau, M. (2001) “The impact of
combinations of building materials and intermittent ventilation on perceived air
quality”, ASHRAE Transactions, Vol. 107(1), pp. 821-835.
Hayashi, T., Ishizu, Y., Kato, S. and Murakami, S. (2002) “CFD analyses on characteristics of
contaminated indoor air ventilation and its application in the evaluation of the effects of
contaminant inhalation by a human occupant”, Building and Environment 37, pp. 219-230.
Heiselberg, P. (1996) “Room air and contaminant distribution in mixing ventilation”,
ASHRAE Transactions, Vol. 102(2), pp. 332-339.
Hinds W.C. (1999) “Aerosol technology: properties, behaviour and measurement of airborne
particles – 2nd ed.”, John Wiley and Sons, ISBN 0-471-19410-7.
Hiwatashi, K., Akabayashi, S., Morikawa, Y. and Sakaguchi, J. (2000) “Numerical study of a
new ventilation tower system for fresh air supply in an air-conditioned room”, Proc. of
Roomvent 2000, 9-12 July, Reading, UK, pp. 565-570.
Holmér, I., Nilsson, M., Bohm, M. and Norén, O. (1999) “Equivalent temperature in vehicles
– conclusions and recommendations for standard”, Proc of 6th International Conference
ATA, 17-19 November, Florence, Italy, pp. 89-94.
Homma, H. and Yakiyama, M. (1988) “Examination of free convection around an occupant’s
body caused by its metabolic heat”, ASHRAE Transactions, Vol. 94(1), pp. 104–124.
Höppe, P. (1981) “Temperature of expired air under varying climatic conditions”, Int. J.
Biometeor, Vol.25, No.2, pp. 127-132.
Hyldgaard, C.E. (1994) “Humans as a source of heat and air pollution”, Proc. of Roomvent
1994, 15-17 June, Krakow, Poland, pp. 413-433.
ISO (1991) “ISO Standard 5167-1: Measurement of fluid flow by means of pressure
differential devices – Part 1: Orifice plates, nozzles and Venturi tubes inserted in circular
cross-section conduits running full”, International Organization for Standardization.
ISO (1993) “Guide to the expression of uncertainty in measurement”, International
Organization for Standardization.
ISO (1994) “ISO Standard 7730: Moderate thermal environments – determination of the PMV
and PPD indices and specification of the conditions for thermal comfort”, International
Organization for Standardization.
Izuhara, I., Kuwahara, R. and Mizutani, K. (2002) “Experimental and numerical studies on
local high efficiency air conditioning system for office buildings”, Proc. of Indoor Air 2002,
30 June-5 July, Monterey, California, pp. 301-306.
Jaakkola, J.J.K., Øie, L., Nafstad, P., Botten, G., Samuelsen, S.O. and Magnus, P. (1999)
“Interior surface materials in the home and the development of bronchial obstruction in
young children in Oslo, Norway”, American Journal of Public Health 89(2), pp. 188-192.
Jaakkola, J.J.K., Verkasalo, P.A. and Jaakkola, N. (2000) “Plastic wall materials in the home
and respiratory health in young children”, American Journal of Public Health 90(5), pp.
797-799.

109
References

Kaczmarczyk, J. (2003) “Human response to personalized ventilation”, PhD thesis,


International Centre for Indoor Environment and Energy, Technical University of
Denmark, ISBN 87-7475-300-2.
Kaczmarczyk, J., Zeng, Q., Melikov A.K. and Fanger P.O. (2002a) “The effect of a
personalized ventilation system on perceived air quality and SBS symptoms”, Proc. of
Indoor Air 2002, 30 June-5 July, Monterey, California, pp. 1042-1047.
Kaczmarczyk, J., Zeng, Q., Melikov, A. and Fanger, P.O. (2002b) “Individual control and
people’s preferences in an experiment with a personalized ventilation system”, Proc. of
Roomvent 2002, 8-11 September, Copenhagen, Denmark, pp. 57-60.
Kaczmarczyk, J., Melikov, A., Bolashikov, Z., Nikolaev, L. and Fanger P.O. (2004) “Thermal
sensation and comfort with five different air terminal devices for personalized
ventilation”, Proc. of Roomvent 2004, 5–8 September, Coimbra, Portugal.
Kaczmarczyk, J., Melikov, A. and Fanger, P.O. (accepted) “Human response to personalized
ventilation and mixing ventilation”, Indoor Air.
Li, Y., Sandberg, M. and Fuchs, L. (1993) “Effects of thermal radiation on airflow with
displacement ventilation: an experimental investigation”, Energy and Buildings 19, pp.
263-274.
Lin, Y.P. and Linden, P.F. (2002) “Modelling an under floor air distribution system”, Proc. of
Roomvent 2002, 8-11 September, Copenhagen, Denmark, pp. 249-252.
Loomans, M.G.L.C (1998) “The measurement and simulation of indoor air flow”, PhD thesis,
Department of Physical Aspects of the Built Environment, Eindhoven University of
Technology, The Netherlands, ISBN 90-6814-085-X.
Loudermilk, K.J. (1999) “Underfloor air distribution solutions for open office applications”,
ASHRAE Transactions, Vol. 105(1), pp. 605-613.
Loudermilk, K.J. (2003) “Temperature control and zoning in underfloor air distribution
systems”, ASHRAE Transactions, Vol. 109(1), pp. 307-314.
Marshall, J.D., Riley, W.J., McKone, T.E. and Nazaroff, W.W. (2003) “Intake fraction of
primary pollutants: motor vehicle emission on the South Coast Air Basin”, Atmospheric
Environment 37, pp. 3455-3468.
Mattsson, M. (1999) “On the efficiency of displacement ventilation with particular reference
to the influence of human physical activity”, Doctoral thesis, Royal Institute of Technology,
Gävle, Sweden, ISBN 91-628-3674-9.
Mattsson, M. (2002) “Vertical distribution of occupant-generated particles in a room with
displacement ventilation”, Proc. of Indoor Air 2002, 30 June-5 July, Monterey, California,
pp. 509-514.
McIntire, DA. (1980) “Indoor Climate”, Applied Science Publishers, ISBN 0-85334-868-5.
Melikov, A.K. and Nielsen, J. B. (1989) “Local thermal discomfort due to draft and vertical
temperature difference in rooms with displacement ventilation”, ASHRAE Transactions,
Vol. 96(2), pp. 1050-1057.
Melikov, A.K. and Langkilde, G. (1990) “Displacement ventilation – airflow in the near
zone”, Proc. of Roomvent 1990, 13-15 June, Oslo, Norway, paper 23.
Melikov, A.K., Langkilde, G. and Derbiszewski, B. (1990) “Airflow characteristics in the
occupied zone of rooms with displacement ventilation”, ASHRAE Transactions, Vol.
96(1), pp. 555-563.
Melikov, A.K. and Zhou, G. (1996) “Air movement at the neck of the human body”, Proc. of
Indoor Air 1996, 21-28 July, Nagoya, Japan, pp. 209-214.
Melikov, A.K., Popiolek Z. and Jørgensen, F.E. (1998) “New method for testing dynamic
characteristics of low-velocity thermal anemometers”, ASHRAE Transactions, Vol. 104(1),
pp. 1490-1506.

110
References

Melikov, A., Kaczmarczyk, J. and Cygan, L. (2000) “Indoor air quality assessment by a
‘breathing’ thermal manikin”, Proc. of Roomvent 2000, 9-12 July, Reading, UK, pp. 101-
106.
Melikov, A.K., Cermak, R. and Majer, M. (2002) “Personalized ventilation: evaluation of
different air terminal devices”, Energy and Buildings 34, pp. 829-836.
Melikov, A.K., Cermak, R., Kovar, O., Forejt, L. (2003) “Impact of airflow interaction on
inhaled air quality and transport of contaminants in rooms with personalized and total
volume ventilation”, Proc. of Healthy Buildings 2003, 4-11 December, Singapore, pp. 592-
597.
Melikov A., Kaczmarczyk., J. and Fanger, P.O. (submitted) “Indoor air quality assessment by
a breathing thermal manikin”, Indoor Air.
Mendell, M.J. (1993) “Non-specific symptoms in office workers: a review and summary of
the epidemiologic literature”, Indoor Air 3, pp. 227-236.
Milton, D.K., Glencross, P.M. and Walters, M.D. (2000) “Risk of sick leave associated with
outdoor air supply rate, humidification, and occupant complaints”, Indoor Air 10, pp.
212-221.
Montgomery, D.C. (2001) “Design and analyses of experiments – 5th ed.”, John Wiley & Sons,
ISBN 0-471-31649-0.
Mundt, E. (1990) “Convection flows above common heat sources in rooms with
displacement ventilation”, Proc. of Roomvent 1990, 13-15 June, Oslo, Norway, paper 38.
Mundt, E. (1995) “Displacement ventilation systems – convection flows and temperature
gradients”, Building and Environment 30, No.1, pp. 129-133.
Mundt, E. (2001) “Non-buoyant pollutant sources and particles in displacement ventilation”,
Building and Environment 36, pp. 829-836.
Murakami, S, Kato, S, Tanaka, T, Choi, D.-H. and Kitazawa, T. (1992) “The influence of
supply and exhaust openings on ventilation efficiency in an air-conditioned room with a
raised floor”, ASHRAE Transactions, Vol. 98(1), pp. 738-755.
Murakami, S., Kato, S. and Zeng J. (1998) “Numerical simulation of contaminant distribution
around a modelled human body: CFD study on computational thermal manikin – Part
II”, ASHRAE Transactions, Vol. 104(2), pp. 226-233.
Murakami, S., Kato, S., Ito, K. and Hayashi, T. (2000) “CFD analyses of indoor chemical
environment and inhaled contaminant by a human body”, Proc. of Cold Climate Heating,
Ventilating and Air-Conditioning, 1-3 November, Sapporo, Japan.
Nielsen, P.V. (1993) “Displacement ventilation – theory and design”, Aalborg University,
ISSN 0902-8002 U9306.
Nielsen, P.V. (1995) “Vertical temperature distribution in a room with displacement
ventilation”, IEA Annex 26: Energy-efficient ventilation of large enclosures, Rome, Italy.
Nilsson, H., Holmér, I., Bohm, M. and Norén, O. (1999) “Definition and theoretical
background of the equivalent temperature”, Proc. of the 6th International Conference ATA,
17-19 November, Florence, Italy.
Papineni, R.S. and Rosenthal, F.S. (1997) “The size distribution of droplets in the exhaled air
of healthy human subjects”, Journal of Aerosol Medicine, Vol. 10, No. 2, pp. 105-116.
Pejtersen, J., Brohus, H., Hyldgaard, C.E., Nielsen, J.B., Valbjørn, O., Hauschildt, P.,
Kjærgaard, S.K. and Wolkoff, P. (2001) “Effect of renovating an office building on
occupants’ comfort and health”, Indoor Air 11(1), pp. 10-25.
Pitchurov, G., Naidenov, K., Melikov, A.K. and Langkilde, G. (2002) “Field study of
occupants thermal comfort in rooms with displacement ventilation”, Proc. of Roomvent
2002, 8-11 September, Copenhagen, Denmark, pp. 479-482.
REHVA (2002) “Guidebook: Displacement ventilation in non-industrial premises”,
Federation of European Heating and Air-conditioning Associations, ISBN 82-594-2369-3.

111
References

Rudnick, S.N. and Milton, D.K. (2003) “Risk of indoor airborne infection transmission
estimated from carbon dioxide concentration”, Indoor Air 13, pp. 237-245.
Skistad, H. (1994) “Displacement ventilation”, Research Studies Press, John Wiley and Sons,
ISBN 0863801471.
Sodec, F. and Craig, R. (1990) “The underfloor air supply system – the European
experience”, ASHRAE Transactions, Vol. 96(2), pp. 690-695.
Spoormarker K.J. (1990) “Low-pressure underfloor HVAC system”, ASHRAE Transactions,
Vol. 96(2), pp. 670-677.
Sundell, J. (1994) “On the association between building ventilation characteristics, some
indoor environmental exposures, some allergic manifestations and subjective symptom
reports”, Indoor Air Suppl. 2, pp. 1-148.
Tanabe, S., Arens, E.A., Bauman, F.S., Zhang, H. and Madsen T.L. (1994) “Evaluating
thermal environments by using a thermal manikin with controlled skin surface
temperature”, ASHRAE Transactions, Vol. 100(1), pp. 39-48.
Tsuzuki K., Arents E.A., Bauman F.S. and Wyon, D.P. (1999) “Individual thermal comfort
control with desk-mounted and floor-mounted task/ambient conditioning (TAC)
systems”, Proc. of Indoor Air 1999, 8-13 August, Edinburgh, Scotland, pp. 368-373.
Wargocki, P., Wyon, D.P., Baik, Y.K., Clausen G. and Fanger, P.O. (1999) “Perceived air
quality, sick building syndrome (SBS) symptoms and productivity in an office with two
different pollution loads”, Indoor Air 9, pp. 165-179.
Wargocki, P., Sundell, J., Bischof, W., Brundrett, G., Fanger, P.O., Gyntelberg, S.O., Hannsen,
S.O., Harrison, P., Pickering, A., Seppänen, O., and Wouters, P. (2002) “Ventilation and
health in non-industrial indoor environments: results from a European Multidisciplinary
Scientific Consensus Meeting (EUROVEN)”, Indoor Air 12, pp. 113-128.
Wargocki, P., Bakó-Biró, Z., Baginska, S., Nakagawa, T., Fanger, P.O., Weschler, C. and
Tanabe, S. (2003) “Sensory emission rates from personal computers and television sets”,
Proc. of Healthy Buildings 2003, 4-11 December 2003, Singapore, pp. 169-175.
Webster, T.L., Bauman, F.S., Reese, J. and Shi, M. (2002a) “Thermal stratification
performance of underfloor air distribution (UFAD) systems”, Proc. of Indoor Air 2002, 30
June-5 July, Monterey, California, pp. 260-265.
Webster, T., Bauman, F. and Reese, J. (2002b) “Underfloor air distribution: thermal
stratification”, ASHRAE Journal, Vol. 44, No. 5 (May), pp. 28-36.
Wolkoff, P. (1995) “Volatile organic compounds – sources, measurements, emissions, and
the impact on indoor air quality”, Indoor Air Suppl. 3, pp. 1-73.
Wolkoff, P., Clausen, P.A. and Nielsen, P.A. (1995) “Application of the field and laboratory
emission cell ‘FLEC’ – performance study, intercomparison study and case study of
damaged linoleum in an office”, Indoor Air 5(3), pp. 196-203.
Yamanaka, T., Satoh, R. and Kotani, H. (2002) “Vertical distribution of contaminant
concentration in rooms with floor-supply displacement ventilation”, Proc. of Roomvent
2002, 8-11 September, Copenhagen, Denmark, pp. 213-216.
Yang, J., Kaczmarczyk, J., Melikov, A. and Fanger, P.O. (2003) “The impact of a personalized
ventilation system on inhaled air quality at different levels of room air temperature”,
Proc. of Healthy Buildings 2003, 4-11 December 2003, Singapore, pp. 345-350.
Zeng, Q., Kaczmarczyk, J., Melikov, A. and Fanger, P.O. (2002) “Perceived air quality and
thermal sensation with a personalized ventilation system”, Proc. of Roomvent 2002, 8-11
September, Copenhagen, Denmark, pp. 61-64.
Zuo, HG., Niu, JL. and Chan, WT. (2002) “Experimental study of facial air supply method
for the reduction of pollutant exposure”, Proc. of Indoor Air 2002, 30 June-5 July,
Monterey, California, pp. 1090-1095.

112
113
Appendix A
Expression of uncertainty

Well-established basic statistical techniques were used for the treatment and analyses of
data. The methods applied have been described elsewhere. In the present study, the
underlying references were Montgomery (2001) and ISO (1993).

Generally, the result of each measurement was characterized by the arithmetic mean of the
observations y accompanied by a statement of the uncertainty of the mean U. The mean y
was the best available estimate of the measured quantity, assuming that the samples were
drawn from an independent population. Violations of the independence (correlation, trends,
etc.) were identified visually using the plot of residuals in time order.

At first, a sample standard uncertainty (level of confidence of 68%) has been calculated for
each quantity. In cases of derived quantities (indexes), the uncertainty was further
propagated. Finally, the uncertainty was expanded in order to increase the level of
confidence to estimated 95%. Section 3.8 summarizes the typical values of uncertainty and
discusses most important sources of uncertainty for each quantity.

Sample standard uncertainty


Each quantity, except for the tracer-gas concentration (see Section 3.6.1), was recorded in
periods several times during each experiment. The periods were identical and adequate in
terms of duration and sampling frequency to estimate the quantity with high accuracy. The
tracer-gas measurement determined the length of an experiment and hence the need for
several measurement periods. The tracer-gas concentrations were recorded continuously in
sequences. The sample standard uncertainty of the mean Ust typically consisted of 3
components:

2 2 2
U st = U meas , max + U stab + U instr (A.1)

where Umeas,max is the maximum uncertainty of measurement (random error);


Ustab is the uncertainty of process stability;
Uinstr is the uncertainty of instrument (calibration).

Table A-1. Standard uncertainty components of measured quantities.


Quantity Umeas Ustab Uinstr
Concentration (inhaled and room)  − 
Inhaled air temperature   
Manikin-based equivalent temp.   
Surface temperature   
Room air temperature −  
Room air velocity −  

115
Appendix A

Table A-1 summarizes the measured quantities and the components of uncertainty used. In
the case of the concentration measurements, the stability of the process was not evaluated
because each measurement was performed in a single sequence (except for the supply and
exhaust air concentration in stages 1 and 2-1, when the whole day measurement was used).
The uncertainty of room air temperature did not include the uncertainty of measurement in
each period, because the information was not available from the instrument. The uncertainty
of velocity measurement was covered by the uncertainty of instrument, provided by the
manufacturer.

Uncertainty of measurement (during each period or sequence)


On the assumption of normality, uncertainty of the mean Umeas,i was expressed using the
standard deviation of the mean. The sample standard deviation s i was computed as a
measure of dispersion of the observed values around their mean y j .

∑(yj =1
i, j − yi )
si
U meas ,i = = (A.2)
m(m − 1) m

where yi , j is the j-th observed value in the i-th period or sequence;


yi is the mean of the observed values in the i-th period or sequence;
si is the sample standard deviation;
m is the number of samples taken during each period or sequence.

The uncertainty of the mean Umeas,i defines an interval having a 68% level of confidence. The
maximum value of uncertainty from the periods (4 or 9 in the case of respectively stages 1
and 2-1 and stage 2-2) was used to calculate the sample standard uncertainty.

Uncertainly of process stability


Each quantity might have changed a little during the whole experiment lasting for 1 day.
The evaluation of process instability was based on the repeated periods of measurements.
The standard deviation of the mean was used:

∑(y
i =1
i − y)
s
U stab = = (A.3)
n(n − 1) n

where y is the grand mean of the observed values;


s is the period standard deviation;
n is the number of measurement periods.

Uncertainty of instrument
Table A-2 presents the uncertainty associated with the instruments used. Three different
approaches were applied:

1. Repeatability of concentration measurement was available from the manufacturer of a


gas analyzer (Section 3.5.3). The repeatability was assumed to have a rectangular
probability distribution, i.e. the measured value lay with equal probability in the interval

116
Appendix A

of ±a. If the differential between the bounds is denoted by 2a, the uncertainty is
calculated as:

a2
U instr = (A.4)
3

2. The uncertainty of the instrument was estimated based on experience in the case of the
inhaled air temperature, manikin-based equivalent temperature and wall surface
temperature. It was assumed that the uncertainty defines an interval having a 95% level
of confidence. Assuming a normal distribution was used, the standard uncertainty (level
of confidence of 68%) was taken as:

U instr ,95
U instr = (A.5)
2

where 2 is the factor corresponding to the 95% level of confidence.

3. Equation provided by the manufacturer was applied for each measurement period in the
case of velocity. The maximum value of uncertainty was taken.

Table A-2. Uncertainty of instrument.


Quantity Uncertainty Comment
Concentration (inhaled and room) 1% of reading* repeatability
Inhaled air temperature 0.2°C** 95% confidence
Manikin-based equivalent temp. 0.2°C** 95% confidence
Wall surface temperature 0.3°C** 95% confidence
Room air temperature 0.3°C* 95% confidence
Room air velocity (0.02· v )+0.022* 95% confidence
* Manufacturer's specification
** Estimate

Uncertainty of derived quantities


In cases of indexes, where a quantity Y was not measured, but was determined from other
quantities through a functional relationship Y = f ( X 1 , X 2 ,..., X N ) , the combined standard
uncertainty Uc,st was estimated. It was given by:

2
N
 ∂f  2
U st ,c = ∑   U st ( xi ) (A.6)
i =1  ∂x i 

where xi is the i-th measured quantity;


Ust(xi) is the standard uncertainty of the quantity.

117
Appendix A

Expanded uncertainty
Expanded uncertainty U is obtained by multiplying the standard uncertainty Ust (or
combined standard uncertainty Ust,c) by a coverage factor k. The value of the coverage factor
k is chosen on the basis of the level of confidence required of the interval covered. If the
probability distribution characterized by y and Ust is normal and the effective degree of
freedom of Ust is of significant size (say greater than 10), the coverage factor k = 2 will
produce an interval y ± kU st corresponding to the level of confidence 95%.

Testing for normal distribution


In order to see if a given set of data followed a normal distribution, a visual examination
using a residual-quantile plot and the numerical Shapiro-Wilk’s w-test was applied. The
residual-quantile plot is a plot of the ordered residuals against the quantiles of a normal
distribution function. If the normal distribution adequately describes the data, the plotted
points fall approximately along a straight line. The plots were constructed using statistical
package S-PLUS 6.

118
Appendix B
Inhaled air concentration with PV, mixing and
displacement ventilation

Figures B-1 and B-2 present the concentration of the floor contaminant, bioeffluents and
exhaled air inhaled by the exposed back manikin with respectively mixing ventilation and
displacement ventilation. The concentrations are compared for the two types of PV and
different scenarios of PV use. It is shown that the use of PV decreased the concentration of
contaminants in the inhaled air as compared to the reference case without PV. The
improvements were different for the two types of terminal, but only small differences were
observed among the three contaminants within the same condition when mixing ventilation
was used. Furthermore, the use of PV at the polluting manikin’s workplace did not affect the
inhaled air quality of the exposed manikin.

1.4
Concentration (C-CS/CE-CS)

1.2 Floor cont.

1 Bioeffluents

0.8
Exhaled air

0.6

0.4
0.2

0
0-0 0-7 0-15 15-0 15-7 15-15 0-7 0-15 15-0 15-7 15-15

RMP VDG
Figure B-1. Mixing ventilation. Concentration of contaminants inhaled by the back manikin.
Different patterns of use are compared: the X-Y labels on the x-axis indicate the PV airflow rates of
the front manikin and the back manikin in L/s per workplace.

1.4
Concentration (C-CS/CE-CS)

1.2 Floor cont.

1 Bioeffluents

0.8
Exhaled air

0.6
0.4

0.2

0
0-0 0-7 0-15 15-0 15-7 15-15 0-7 0-15 15-0 15-7 15-15

RMP VDG
Figure B-2. Displacement ventilation. Concentration of contaminants inhaled by the back manikin.
Different patterns of use are compared: the X-Y labels on the x-axis indicate the PV airflow rates of
the front manikin and the back manikin in L/s per workplace.

119
Appendix C
Inhaled air temperature with PV, mixing and
displacement ventilation

Figures C-1 and C-2 present the inhaled air temperature for the front manikin and the back
manikin, respectively. The temperatures are compared for the two types of PV coupled with
mixing ventilation and displacement ventilation at different combinations of PV supply
rates. It is shown that the use of PV at one workplace did not affect the inhaled air
temperature of the occupant at another workplace by more than 0.3°C, regardless of the
total-volume ventilation system. The impact of the PV airflow rate on the inhaled air
temperature is apparent, as is the ability of the displacement ventilation to provide a lower
inhaled air temperature than mixing ventilation (both with and without PV).

28
27 Mixing
Temperature (°C)

26 Displacement
25
24
23
22
21
20
0-0 0-7 0-15 15-0 15-7 15-15 0-7 0-15 15-0 15-7 15-15

RMP VDG
Figure C-1. Inhaled air temperature of the front manikin. Different patterns of use are compared: the
X-Y labels on the x-axis indicate the PV airflow rates of the front manikin and the back manikin in
L/s per workplace.

28
27 Mixing
Temperature (°C)

26 Displacement
25
24
23
22
21
20
0-0 0-7 0-15 15-0 15-7 15-15 0-7 0-15 15-0 15-7 15-15

RMP VDG
Figure C-2. Inhaled air temperature of the back manikin. Different patterns of use are compared: the
X-Y labels on the x-axis indicate the PV airflow rates of the front manikin and the back manikin in
L/s per workplace.

121
Appendix D
Whole-body manikin-based equivalent temperature

Figures D-1 and D-2 present the whole-body manikin-based equivalent temperatures (ET)
for PV combined with respectively mixing and displacement ventilation at different rates of
personalized air. Although different in magnitude (due to difference between the
workplaces), the temperature characteristics are similar in the two figures. As expected, an
increase in the PV airflow rate increased cooling and hence decreased the ET. It is shown
that the use of PV by one manikin did not affect the ET of the other manikin.

27
Front
26 manikin
Back
ET (°C)

25
manikin
24

23

22
0-0 0-7 0-15 15-0 15-7 15-15 0-7 0-15 15-0 15-7 15-15

RMP VDG

Figure D-1. Mixing ventilation. Whole body manikin-based equivalent temperature. Different
patterns of use are compared: the X-Y labels on the x-axis indicate the PV airflow rates of the front
manikin and the back manikin in L/s per workplace.

27
Front
26 manikin
Back
ET (°C)

25
manikin
24

23

22
0-0 0-7 0-15 15-0 15-7 15-15 0-7 0-15 15-0 15-7 15-15

RMP VDG

Figure D-2. Displacement ventilation. Whole body manikin-based equivalent temperature. Different
patterns of use are compared: the X-Y labels on the x-axis indicate the PV airflow rates of the front
manikin and the back manikin in L/s per workplace.

123
Appendix E
Inhaled air concentration with personalized and
underfloor ventilation

Figures E-1 to E-4 present the inhaled air concentration of the floor contaminant, human
bioeffluents and exhaled air with underfloor air distribution for different discharge patterns
and different supply airflow rates. The concentrations are compared for the two types of PV
performing under different patterns of use. The X-Y labels on the x-axis indicate the PV
airflow rates of the front manikin and the back manikin in L/s per workplace.

1.4
Concentration (C-CS/CE-CS)

Floor cont. inh.


1.2
to front man.
1 Floor cont. inh.
0.8 to back man.

0.6 Bioeffluents inh.


to back man.
0.4
Exhaled air inh.
0.2 to back man.

0
0-0 0-15 15-15 15-0 0-15 15-15 15-0

RMP VDG
Figure E-1. Vertical discharge of UFAD, stage 2-1 (VAV strategy). Concentration of contaminants
inhaled by the front manikin and the back manikin.

1.4
Concentration (C-CS/CE-CS)

Floor cont. inh.


1.2
to front man.
1 Floor cont. inh.
0.8 to back man.

0.6 Bioeffluents inh.


to back man.
0.4
Exhaled air inh.
0.2 to back man.

0
0-0 0-15 15-15 15-0 0-15 15-15 15-0

RMP VDG
Figure E-2. Horizontal discharge of UFAD, stage 2-1 (VAV strategy). Concentration of
contaminants inhaled by the front manikin and the back manikin.

125
Appendix E

Concentration (C-CS/CE-CS) 1.4


Floor cont. inh.
1.2
to front man.
1 Floor cont. inh.
0.8 to back man.

0.6 Exhaled air inh.


to front man.
0.4
Exhaled air inh.
0.2 to back man.
NA
0
0-0 0-15 15-15 15-0 0-15 15-15 15-0

RMP VDG
Figure E-3. Vertical discharge of UFAD, long throw (80 L/s), stage 2-2 (CAV strategy).
Concentration of contaminants inhaled by the front manikin and the back manikin.

1.4
Concentration (C-CS/CE-CS)

Floor cont. inh.


1.2
to front man.
1
Floor cont. inh.
0.8 to back man.

0.6 Exhaled air inh.


to front man.
0.4
Exhaled air inh.
0.2 to back man.

0
0-0 0-15 15-15 15-0 0-15 15-15 15-0

RMP VDG
Figure E-4. Vertical discharge of UFAD, short throw (50 L/s), stage 2-2 (CAV strategy).
Concentration of contaminants inhaled by the front manikin and the back manikin.

126
Appendix F
Vertical distribution of active contaminants with
underfloor ventilation

Yamanaka et al. (2002) proposed a model that aims at predicting the vertical distribution of
active contaminants in rooms with underfloor ventilation. The distribution is three-zonal
(Figure F-1). The model makes it possible to predict the concentration of contaminants in the
lower zone as well as the concentration of contaminants in the interface layer.

cU
Height
VU cU VMT above Maximum
floor throw
VPE cPE
VINT Interface
VUE VL cL height

V
cL Concentration
Figure F-1. Air distribution model for underfloor ventilation.

Although not explained in detail in their article, the model is obviously based on a mass
balance for the interface layer. If the vertical flow from underfloor ventilation penetrates the
height of stratification (see Figure 1.3), air and contaminants from the upper zone are forced
to flow downward. Part of the down flow is entrained by the plumes above occupants and
heated objects. The rest of the down flow brings contaminants to the lower zone. The mass
balance for the interface layer can be written as:

VU cU = VPE c PE + VL c L (F.1)

where VU is the down airflow rate from the upper zone,


cU is the concentration of contaminants in the upper zone,
VPE is the airflow rate entrained by thermal plumes,
cPE is the concentration of contaminants in the interface layer, entrained by
thermal plumes,
VL is the down airflow rate to the lower zone,
cL is the concentration of contaminants in the lower zone.

According to Yamanaka et al. (2002), the top of the interface layer is located at the same
height as maximum throw of vertical flow from underfloor ventilation. They defined the
maximum throw height as the height where velocity decreases to zero (instead of the height
where velocity decreases to 0.25 m/s, as used in Chapter 5). Moreover, they defined the
down flow rate as the surplus flow rate of the plume flow at the top of the interface layer:

127
Appendix F

VU = VMT − V (F.2)

where VMT is the airflow rate in the thermal plume at the height of maximum throw,
V is the airflow rate of underfloor ventilation,

The airflow rate entrained by thermal plumes (VPE) is equal to the airflow rate in the thermal
plumes at the height of maximum throw (VMT) minus the airflow rate in the thermal plumes
at the height of interface (VINT). The down airflow rate to the lower zone was assumed to be
equal to the airflow rate in the thermal plume at the height of the interface, i.e. VL = VINT.
Entrainment of air by the vertical flow of underfloor ventilation (VUE) was not considered in
the original model. The concentration of contaminants in the upper zone was assumed to be
1 (normalized for the exhaust concentration), i.e. cU = 1. Yamanaka et al. (2002) assumed that
the concentration of the airflow from the interface layer to thermal plumes is equal to the
average value of the upper zone concentration and the lower zone concentration. Hence, the
concentration of contaminants in the lower zone can be expressed from Equation F.1 as:

2(VMT − V ) − VMT + VINT


cL = (F.3)
VMT + VINT

Yamanaka et al. (2002) validated the model using a scaled experimental room. The
ventilation air was supplied from the floor by means of 16 round openings. Two ventilation
airflow rates and three throw heights (achieved by changing the diameter of the openings)
were studied. Four person simulators were arranged on the floor and a tracer-gas was
released at the top of the simulators. The vertical distribution of the contaminants was
measured. The widths of the interface layer (W, in metres), which is needed for the
definitions of the interface height, were determined experimentally:

W = 0.000956 VU (F.4)

where VU is the down airflow rate from the upper zone in m3/h.

Application
The model by Yamanaka et al. (2002) was applied in the present study in order to determine
the vertical distribution of human-produced contaminants in the case of underfloor
ventilation alone. The comparison was made for two airflow rates of underfloor ventilation
(vertical discharge direction), namely 80 and 50 L/s. The maximum throw at the two rates
was estimated respectively at 1.5 m and 0.8 m, based on Figure 5.1. Table F-3 presents the
volumetric flows in thermal plumes at the height of maximum throw and the height of the
interface. The flows were estimated based on charts presented by Nielsen (1993), reprinted
also in REHVA (2003) guidebook. Equation F.4 was used to calculate the width of the
interface: 0.35 and 0 m for the two cases, respectively.

Figure F-2 compares the predicted contaminant concentration with the contaminant
concentration of exhaled air (both manikins, averaged) measured near the two workplaces
(positions A and B, averaged) in the present study. The two distributions are in good
agreement. However, more full-scale measurements are needed in order to validate the
model by Yamanaka et al. (2002) in real settings with certainty.

128
Appendix F

Table F-3. Convection volume flow above occupants and heated objects per workplace in L/s.
Elevation of the objects above floor is indicated in parenthesis.
Heat source Long vertical throw Short vertical throw
Maximum Interface Maximum Interface
throw height height throw height height
(1.5 m) (1.15 m) (0.8 m) (0.8 m)
Person 30 20 15 15
PC tower (0.3 m) 30 20 15 15
PC monitor (0.8 m) 25 10 0 0
Desk lamp (1 m) 5 0 0 0
Total 90 50 30 30

2.5 2.5
Measured Measured
Height above floor (m)

2 Calculated Height above floor (m) 2 Calculated

1.5 1.5

1 1

0.5 0.5
Long throw Short throw
0 0
0 0.5 1 1.5 0 0.5 1 1.5
Concentration (C-CS/CE-CS) Concentration (C-CS/CE-CS)
Figure F-2. Comparison of concentration profiles measured (see Figure 5.7) and calculated based on
Yamanaka et al. (2002).

The model can be improved by considering the ability of the vertical flows of underfloor
ventilation to transport air from the lower zone to the interface layer. If included, the down
airflow rate to the lower zone can be assumed to be equal to the airflow rate in the thermal
plume at the height of the interface plus the airflow rate entrained by the vertical flow of
underfloor ventilation in the lower zone, i.e. VL = VINT + VUE . However, the predictions
using the original model (airflow from underfloor ventilation not considered) agreed well
with the results of measurement in the present study. The reason may be the fact that
relatively small amounts of air were transported by the vertical flows of underfloor
ventilation from the lower zone to the interface layer, as indicated by low air velocity at the
interface height (Figure 5.3). In other situations, especially when the vertical throw is long,
the impact of the vertical flows of underfloor ventilation on the contaminant distribution
may be important.

129
Appendix G
Risk of airborne infection transmission

The Wells-Riley equation is commonly used to model the risk of airborne transmission of
infectious diseases. Because air exhaled by people is associated with a release of infectious
agents, Rudnick and Milton (2003) derived an alternative equation that determines the risk
of transmission by using CO2 concentration as a marker for exhaled-breath exposure. The
normalized concentration of exhaled air contaminants (= 1/ventilation effectiveness) was
introduced in the model of Rudnick and Milton (2003) in order to quantify the impact of
various ventilation systems on the risk of infections.

The volume fraction of inhaled air that is exhaled breath is determined as:

Ve C − CS
f = C (− ) = (G.1)
V Ca

where Ve is the equivalent volume of exhaled breath contained in indoor air, m3;
V is the volume of the shared air space, m3;
C(-) is the normalized concentration, see Section 3.7.1;
C is the volume fraction of the exhaled air tracer in inhaled air;
CS is the volume fraction of the tracer in supply air,
Ca is the volume fraction of the tracer added to the exhaled breath during
breathing.

Rudnick and Milton (2003) used CO2 produced by people as a marker of exhaled breath. In
the present study the exhaled air was marked with a constant dose of SF6 for the front
manikin and N2O for the back manikin (stage 2-2). Because every person in a shared space
produces CO2, the volumetric fraction of inhaled air that was exhaled by infectors had to be
determined:

fI
f′= (G.2)
n

where I is the number of infectors;


n is the number of persons in a ventilated space.

In the present study there was only a single infector (the manikins used different tracers),
and nobody in the room contributed to the tracer-gas concentration in the room, hence
f′= f .

Example:
A room with mixing air distribution (i.e. normalized concentration in inhaled air equal to 1)
is ventilated with 80 L/s of clean air. The pulmonary ventilation of occupants is 6 L/min (0.1
L/s). Alternatively, exhaled air from the front manikin contains about 1600 ppm SF6, while
the concentration of SF6 inhaled by the back manikin is 2 ppm (measured).

131
Appendix G

0 .1 2−0
f = f′= 1= = 0.00125 (G.3)
80 1600

The calculation reveals that the air inhaled by the back manikin contains 0.125% of air
exhaled from the front manikin.

The basic reproductive number (R0) is the number of secondary infections that arise when a
single infectious case is introduced into a population where everyone is susceptible. An
infectious agent can spread in a given population, if R0 > 1. The larger the value of R0 the
more likely is the infection to reproduce rapidly in the form of an epidemic. The
reproductive number in a building environment (RA0) can be derived (see Rudnick and
Milton, 2003) as:

R A0 = (n − 1)[1 − exp(− f ′qt )] (G.4)

where q is the quantum generation rate by an infected person, quanta/hour;


t is the total exposure time, hour.

Rudnick and Milton (2003) stress that q represents the generation rate of infectious doses,
not organisms or infectious particles; hence, it is the average infectious source strengths of
infected individuals. The quantum generation rate depends on the infectious agent. Rudnick
and Milton (2003) assessed the risk of 3 infections with a different q value: measles: q = 570
quanta/hour, influenza: q = 100 quanta/hour, rhinovirus: q = 4 quanta/hour. The same
estimates are used in the following analyses.

There are several factors that determine the reproductive number: outdoor ventilation rate,
effectiveness of the ventilation system, quantum generation rate (infection), number of
occupants in a shared space and total exposure time. Figure G-1 demonstrates the risk of
infections in a typical open plan office. It shows the reproductive number as a function of the
normalized concentration of an agent in inhaled air (= 1/ventilation effectiveness) for 3
different infections. There are 30 persons in the room (chosen arbitrarily) and the outdoor
ventilation rate is 10 L/s per person, which corresponds to the typical requirements of
present standards (CEN, 1998; ASHRAE, 1994). The time of exposure is 8 hours. The figure
shows that in the case of mixing ventilation (i.e. normalized concentration of 1) there is a
likelihood that 23 out of 30 persons contract measles, if a single infector were present. In the
same setting 7 out of 30 persons may contract influenza. Rhinovirus would not spread (RA0 <
1). The risk of infections decreases with the decrease in the normalized concentration, i.e. as
the efficiency of ventilation increases.

It is obvious that the number of secondary infections depends on the ventilation rate, i.e. on
the diluting of the infectious dose generated. Figure G-2 compares the reproductive number
for three supply rates of outdoor air in an example of influenza. The number of occupants is
again 30 and the exposure time is 8 hours. It is shown that a ventilation effectiveness as high
as 10 would be required in order to prevent the spread of influenza at a ventilation rate of 10
L/s per person, while at a rate of 40 L/s per person a ventilation effectiveness of just 2
would be sufficient. Wargocki et al. (2002) concluded, based on scientific data, that outdoor
air supply rates above 25 L/s per person are likely to decrease the risk of health and comfort
problems (SBS symptoms) and increase productivity. Figure G-2 demonstrates that a
ventilation effectiveness of 3 would be needed in order to prevent the spread of influenza in
such a case.

132
Appendix G

Reproductive number, RA0 25


Measles
20 Influenza
Rhinovirus
15

10

0
0 0.2 0.4 0.6 0.8 1
Concentration (C-CS/CE-CS)

Figure G-1. Reproductive number vs. dimensionless concentration of measles, influenza or


rhinovirus.

8
Reproductive number, RA0

7 10 L/s per person


6 25 L/s per person
40 L/s per person
5
4
3
2
1
0
0 0.2 0.4 0.6 0.8 1
Concentration (C-CS/CE-CS)

Figure G-2. Reproductive number vs. dimensionless concentration of influenza at three levels of
outdoor ventilation rate for influenza.

133
Appendix H
Intake fraction

Intake fraction (iF) has been used in the environmental health literature to express the
source-to-intake relationship. The intake fraction is defined as the integrated incremental
intake of a pollutant, summed over all exposed individuals, and occurring at any time,
released from a specific source, per unit of pollutant emitted (Bennett et al., 2002). iF can be
expressed as:

∫ nQC (t )dt
T1
iF = T2
(H.1)

∫ E (t )dt
T1

where T1, T2 are the starting and ending time of the emission, s,
n is the number of occupants exposed,
Q is the breathing rate of an occupant, m3/s,
C is the inhaled air concentration from a specific source, g/m3,
E is the emission of the source, g/s.

For indoor environments, the inhaled air concentration from a specific source (C) can be
determined based on the ventilation rate and the emission of the source. In rooms with a
non-uniform distribution of contaminants, the ventilation effectiveness (or the normalized
concentration) must be considered. Equation G.1 became:

E
nQ C (− )
V nQC (− )
iF = = (H.2)
E V

where V is the room ventilation rate, m3/s,


C(-) is the normalized concentration (= 1/ventilation effectiveness).

It is shown that iF became independent of the emission source and its strengths. Assuming a
constant breathing rate, the intake fraction depends entirely on the ventilation rate, the
ventilation effectiveness (i.e. normalized concentration) and the number of occupants in a
room. In addition, a whole-day exposure under a steady-state condition was assumed in
Equation H.2 for simplification.

In the present study (Figure 7.2), the breathing rate (Q) of an occupant performing light
office work was assumed to be 6 L/min (8.6 m3/day). This is lower than 20 m3/day,
typically used in environmental health literature.

135

Potrebbero piacerti anche