ArticlePDF Available

Abstract and Figures

Mould is not a new phenomenon in buildings as the above quote from the Bible indicates. One of the main health issues often related to affordable warmth is that of mould growth and dampness in dwellings. This chapter reviews the current results of research into the causes of mould growth, proposes a mould index for dwellings and investigates the impact of fuel poverty on the occurrence of mould growth.
Content may be subject to copyright.
2.2
Mould Index
Tadj Oreszczyn, University College London
Stephen Pretlove, South Bank University
“then he who owns the house shall come and tell the priest, ‘There seems to
me to be some sort of disease in my house.’ Then the priest shall command
that they empty the house before the priest goes to examine the disease, least
all that is in the house be declared unclean; and afterward the priest shall go
in to see the house. And he shall examine the disease; and if the disease is in
the walls of the house with greenish or reddish spots, and if it appears to be
deeper than the surface, then the priest shall go out of the house to the door
of the house, and shut up the house seven days. And the priest shall come
again on the seventh day, and look; and if the disease has spread in the walls
of the house, then the priest shall command that they take out the stones in
which is the disease and throw them into an unclean place outside the city;
and he shall cause the inside of the house to be scraped round about, and the
plaster that they scrape off they shall pour into an unclean place outside the
city; then they shall take other stones and put them in the place of those
stones, and he shall take other plaster and plaster the house.”
LEVITICUS 14 35-43
Mould is not a new phenomenon in buildings as the above quote from the
Bible indicates. One of the main health issues often related to affordable
warmth is that of mould growth and dampness in dwellings. This chapter
reviews the current results of research into the causes of mould growth,
proposes a mould index for dwellings and investigates the impact of fuel
poverty on the occurrence of mould growth.
1. The problem
Approximately 20% of all dwellings in England suffer from mould growth
and dampness to some degree. This number of incidences of mould growth
and dampness in private rented dwellings, so often occupied by the fuel poor,
is significantly higher than in any other sector (DOE 1996).
There have been many studies that suggest a link between health and
dampness and mould (IEA 1991). In fact one could argue that the causal link
is perhaps stronger than between low internal temperatures or fuel poverty
and health. However, it is always difficult to identify a causal relationship
because of the many confounding factors.
Mould growth is not the only potential health hazard resulting from
dampness and high internal relative humidities. The House-Dust Mite (HDM)
which thrives in high relative humidities is considered to be an important
causal agent not only for asthma, but also for other allergic disease, such as
atopic dermatitis, rhinitis and keratoconjunctivitis (Cunningham 1996). The
MRC Institute for Environment and Health conclude that "There is clear
evidence that antigen derived largely from mite faeces is one of the major
causes of allergic sensitisation in the UK" (Humfrey 1996).
2. Conditions for mould growth and dust mites
There are many different types of mould species, the most common in UK
dwellings include the genera Aspergillus, Penecillium and Cladosporium.
There are other genera commonly found in dwellings and they all thrive
under slightly different environmental conditions. However, all moulds
require the following to grow.
Mould spores: these normally enter the building with the air from outside
which normally contains several hundred spores per cubic meter although
the concentration increases in summer when external concentrations are
considerably higher than internal concentrations, even if the property has
mould growth.
Nutrients: the slightest layer of grease found on all building surfaces is
adequate for mould to grow.
Temperature: different mould species thrive under different temperatures.
Most moulds will survive in the range 0°C to 40°C, found in most
buildings. Moulds will normally thrive better in warmer conditions.
However, over the range of temperatures normally found in dwellings,
10°C to 20°C, there is little change in activity.
Moisture: The availability of moisture is normally the critical factor
determining whether mould will grow in a dwelling. Moulds extract
moisture from the substrate they are growing on which in turn extracts
moisture from the surrounding air. Therefore, both the surface properties
and local relative humidity are important factors. Mould can grow on
hygroscopic materials such as wood and leather at a relative humidity as
low as 70% whereas on surfaces which do not absorb moisture such as
glass and ceramic tiles require water to condense on the surface which
will occur at a relative humidity of 100%. For common wall coverings
such as painted plaster or wallpaper the critical surface relative humidity
for mould growth is 80%. If a relative humidity of 80% is maintained for
a period of several weeks then mould will grow. However, conditions
within dwellings are rarely stable, either because the external climate is
varying, the heating system is being turned on and off or moisture is
being introduced into the dwelling through the occupants breathing,
cooking or washing. Few detailed studies have been undertaken into the
impact of transient conditions on mould growth and those that have,
indicate that mould will grow if the relative humidity is above 80% for
50% of the time (Adan 1994). Analysis of monitored data from dwelling
living and bedrooms indicate that this corresponds to a mean relative
humidity of 80% (Oreszczyn & Pretlove 1998).
Proliferation of the house dust mite (HDM) can be reduced by removing the
source of food (human skin scale) by microporous barriers, killing mites with
acaricide sprays, heat treatments, freezing or dehydrating mites. Controlling
the relative humidity and temperature is the only long term cost effective
method to control dust mite proliferation prior to sensitisation. The dominant
HDM in the UK, Dermatophagoides pteronyssinus, rapidly increases its
feeding, defecating, mating and egg production above 73% relative humidity
at temperatures above 25°C. The HDM depends on mould to breakdown skin
scales and make them digestible. However, if the relative humidity is above
85% the mould activity results in scales so decomposed as to make them
inedible to the mites. Adult mites dehydrate and can survive no longer than 6-
11 days at a relative humidity less than 50%. As temperature falls, egg output
slows and egg-to-adult development time rapidly extends from 34 days at
23°C to 123 days at 16°C (at 75% relative humidity). This is one of the
reasons why central heating has been suggested as a possible cause of
asthma. However, the evidence for this is inconclusive and whereas the
temperature associated with central heating may encourage mite populations
to expand, central heating should substantially reduce the relative humidity
therefore reducing mite numbers. Temperatures have to remain below
freezing for some time in order to kill adult mites. Although many texts refer
to maintaining an absolute humidity below 7g/kg, as proposed by Korsgaard
(1983), this rule of thumb only applies to well-heated dwellings. For room
temperatures below 20°C the relative humidity of surfaces with soft
furnishings should be kept below 75% with occasional periods of a week or
two below 50%. This is most easily achievable in dwellings during cold
spells when the external air is very dry.
The above clearly identifies relative humidity as the key parameter for
determining if mould or dust mites will survive. As relative humidity is a
function of both the moisture and the temperature of the air the next sections
identify the key factors that impact on these, both in general terms and
specifically at the locations where mould and mites live.
3. Micro environmental conditions.
Moulds and dust mites live on surfaces and within soft furnishings
respectively and hence it is the microenvironments that surround these
locations that are of importance and not just the average airspace conditions
within a dwelling. This is particularly the case in older uninsulated properties
so often occupied by the fuel poor.
A relative humidity of above 70% is often quoted (BSI 1989) as the critical
condition for mould growth. This figure refers to the average conditions of
the air that is typically found in dwellings where mould occurs. However,
mould and house dust mites grow on surfaces and the local surface conditions
can be significantly different from the air temperature. Since relative
humidity is a function of both temperature and the moisture content of the air,
the local relative humidity close to the surface of a colder wall will be higher
than the relative humidity in the bulk of the room. This is particularly the
case in dwellings that are only partially heated, uninsulated, or where thermal
bridges occur, such as where insulation is missing or where there is two or
three-dimensional heat loss (Oreszczyn 1992). In order to keep the surface
relative humidity below 80%, and therefore avoid mould growth, on an
uninsulated brick wall the critical airspace relative humidity can be as low as
65%. For an well-insulated cavity wall the critical airspace relative humidity
can be 75% and on a typical cold bridge it can be as low as 60%. For the
above reasons there may be cases when the average room conditions are
below the critical levels when either mould or mites should flourish but
where there are specific microenvironments where moulds or mites can and
do thrive.
4. Moisture production and movement in dwellings
The amount of moisture in the air within a room is dependent on the moisture
produced within a space and the moisture transported into and out of the
dwelling predominately by ventilation.
Moisture production
The amount of moisture produced and released into the air within a dwelling
can range between 1 and more than 20 litres per day, depending on the
number of occupants and their moisture production activities. Figure 1 shows
a typical moisture production profile for a family of four. The largest
moisture source is due to the perspiration and exhaled air of the occupants,
followed by cooking and then bathing. However, personal habits can
radically affect this. For example, microwave cooking of ready-prepared
meals can substantially reduce the moisture associated with cooking. Also,
occupants who spend all day inside the dwelling, such as retired couples, will
result in increased occupant moisture production when compared to working
couples. Dwellings should be designed to cope with a typical range of
moisture production in a dwelling of a given size. However, any dwelling can
experience mould growth if excessive moisture is generated. Dwellings
should be designed to cope with some internal drying of clothes and if the
dwelling is a flat without a secure garden or laundry facility it is likely that a
substantial proportion of drying will be carried out indoors on radiators.
Figure 1 - Typical moisture production for a family of four (kg/day)
Moisture transport
The predominant mechanism for moisture transfer in dwellings is ventilation.
Figure 2 summarises the typical daily moisture flows in a dwelling. The
incoming air has moisture associated with it and additional moisture is added
to this air due to the moisture production detailed above. This moisture is
then transported through the fabric of the dwelling by diffusion and by the
extract air. This air has greater moisture content than the incoming air so
increasing the ventilation rate reduces the internal moisture content. The
moisture transfer through the fabric by diffusion is negligible in terms of the
total quantity directly carried by the air. If this moisture condenses as it
diffuses through the fabric then interstitial condensation may arise and the
fabric may be damaged. Such interstitial condensation can occur if a wall is
drylined with insulation but no vapour barrier is introduced.
3.95
2.20
0.50
0.88
0.45
1.25
1.00
0.20
Occupants
Cooking
Washing clothes
Washing floors & surfaces
Washing dishes
Bathing / washing
Plants
Total = 10.43 kg/day
Figure 2 - Typical daily moisture flows in a dwelling
Up to 200 litres of moisture may be stored in the fabric of a typical dwelling.
As moisture is released into the space some will be absorbed into the fabric
and then when the moisture production is stopped the moisture may then be
re-emitted into the internal air. The extent of moisture absorption and
desorption depends on the surface materials inside the building. An example
of a highly absorbent environment would be a wood panelled room with very
hygroscopic materials within such as a library. The other extreme could be a
fully tiled bathroom, which contains minimal hygroscopic material.
However, although moisture storage in the fabric plays an important role in
short term fluctuations, what is absorbed is generally re-emitted and so this
phenomena can be ignored if one is concerned with long term averages, as is
the case with mould growth. The one main exception to this is where the
fabric of a new dwelling is still drying out following construction. Evidence
collected from a variety of new dwellings suggests that significantly elevated
internal relative humidity is experienced for about 6 months to a year after
construction is completed which can lead to mould growth.
5. Internal temperature
The internal relative humidity is not only dependent on the internal moisture
but also on the internal temperature. The lower the temperature for given
moisture content the higher the relative humidity. Therefore as dwellings
become hard to heat and occupants cannot afford to heat them the relative
humidity rises within the dwellings. What’s more although ventilation may
reduce the moisture content of the air it will also reduce the temperature in a
dwelling occupied by the fuel poor. In these situations, when the ventilation
rate increases beyond a certain level the temperature drops and so the relative
humidity rises. These two factors often compete against each other. Therefore
200kg stored in building fabric
30 to 60 kg/day
in incoming air
6 kg/day swing
with the building
fabric
Occupants 3
to 20 kg/day
0.5 kg/day in
diffusion losses
30 to 80 kg/day
in exhausted
ventilatio n air
whereas a strategy of increased ventilation to eliminate mould growth may be
appropriate in a fuel rich property it may be less so for a fuel poor dwelling.
This is indicated in Figure 3 below. In addition to the ventilation rate, the
temperature will be affected by the type of heating system, its control, cost of
operation and the levels of fabric insulation.
Figure 3 - The impact of increased ventilation on internal relative humidity
6. Relative humidity modelling
Because the internal surface relative humidity is dependent on both the
moisture and the temperature, the relative humidity is dependent on many
inter-related factors including the following:
Fabric Insulation
Occupants ability to pay fuel bills
Occupants moisture generating activities
Hours of occupation
Number of occupants
Heating system, efficiency, control and distribution
Ventilation
External climate
In determining the internal surface relative humidity, many of the above
factors interact in a relatively complex manner. As a result of this complexity
a computer model called Condensation Targeter II (Oreszczyn & Pretlove
1998) has been created in order to gain an insight into the potential risk of
40
50
60
70
80
90
100
00.5 11.5 22.5 3
Air ch an ges p er h our
Internal RH
"Fuel poor"
"Fuel rich"
Crit ic al Inte rn al RH
for Mould Growth = 70%
mould growth in a particular dwelling with a standard occupant. This model
utilises the thermal model BREDEM-8 to predict the internal temperature.
Added to this monthly temperature calculation is a simple algorithm for
calculating the surface temperature and a steady state moisture model
(Loudon 1971). Predictions of internal relative humidities for 36 dwellings,
mostly modern, have been compared with measured relative humidities and
shown to agree to within +/- 10% for 95% of cases as indicated in Figure 4.
Figure 4 - Condensation Targeter II monthly modelled versus monitored
relative humidity (Oct-May) for 36 dwellings
7. Sensitivity
In order to determine the impact that different factors have on the risk of
mould growth, a sensitivity study using Condensation Targeter II has been
undertaken. Figure 5 shows the impact of changing fabric, services and
occupancy factors on the average air relative humidity within an uninsulated
top storey flat.
0.00
20.00
40.00
60.00
80.00
100.00
0.00 20.00 40.00 60.00 80.00 100.00
Monitored RH (%)
Modelled RH (%)
+10% RH
-10% R H
Figure 5 - Sensitivity study for uninsulated top-storey flat with two exposed
walls and roof where fuel cost is no object
This flat is modelled as having an occupant who can afford to heat the
property to 21°C for an average 10 hours a day and the standard case has a
predicted relative humidity of 71%. The first bar shows the impact of
changing the heating pattern. If heating is reduced to only 5 hours a day then
the relative humidity increases to 92%. Alternatively if the flat is heated for a
total of 16 hours a day then the relative humidity drops to 58%. The different
interventions are ordered from left to right as roughly having less
significance. Figure 6 shows the same dwelling, but this time the tenants are
assumed to have a limited fuel budget of £ 60 per week. In this case
insulation has as significant an impact as ventilation.
Figure 6 - Sensitivity study for uninsulated top-storey flat with two exposed
walls and roof where fuel cost limited to £60/week
50.00
55.00
60.00
65.00
70.00
75.00
80.00
85.00
90.00
95.00
100.00
Heating pattern
Occupancy
Ventilation rate
Demand
temperature
Geographical
location
Insulation
Clothes drying
indoors
Unvented gas
heating
Number of
chimneys
Window type
Location of
kitchen
Input parameter
Airspace Relative Humidity (%)
Worst
Best
Never
5 hours/day
Never
7 loads/week
South West
(January)
East Scotland
(January)
Zone 1
Zone 2
Unopenable
All very loose
None
2 Chimneys
19 C
23 C
Highest
1.15 ach
Lowest
0.56 ach
Two
Six
Extended
16 hours/day
Evenings only
5 hours/day
Standard dwelling airspace Relative Humidity = 71%
50.00
55.00
60.00
65.00
70.00
75.00
80.00
85.00
90.00
95.00
100.00
Heating pattern
Occupancy
Ventilation rate
Demand
temperature
Geographical
location
Insulation
Clothes drying
indoors
Unvented gas
heating
Number of
chimneys
Window type
Location of
kitchen
Input parameter
Airspace Relative Humidity (%)
Worst
Best
Never
5 hours/day
Never
7 loads/week
South West
(January)
East Scotland
(January)
Zone 1
Zone 2
Unopenable
All very loose
None
2 Chimneys
Highest
1.15 ach
Lowest
0.56 ach
Two
Six
Extended
16 hours/day
Evenings only
5 hours/day
Standard dwelling airspace Relative Humidity = 75%
8. Mould Index
Chapter (?) proposes the introduction of an Affordable Warmth Index (AWI).
A similar index has been developed called the Mould Index (MI). The Mould
Index indicates the risk of mould occurring on the coldest surfaces in a
dwelling. If the average surface relative humidity of the coldest surface
equals 80% in any month then the MI will equal zero, whilst if the average
surface relative humidity equals 50% in any month then the MI is given a
value of 100. Between these two criteria the MI is assumed to be linear and
can extend below and beyond the 0 and 100 MI ratings. A high value on the
AWI scale indicates affordable warmth in a dwelling and a high value on the
MI scale indicates avoidance of mould growth.
Using the computer model described earlier it is possible to examine the
impact that different interventions have on the AWI and MI. It is also
possible to model dwellings where fuel cost is limited and where it is
unlimited which has an impact on the MI values. As the level of insulation is
improved in a poorly insulated dwelling the AWI and the MI increases.
These are indicated in Figure 7 below.
Figure 7 - Effect of insulation on AWI and MI
As expected, the fuel poor MI values, for situations where there is limited
fuel expenditure, are lower than the fuel rich MI values where fuel cost is
unlimited. Although the trend is similar, as the level of insulation in a
dwelling reduces mould growth is far more likely in dwellings of the fuel
poor.
The impact of ventilation rate has also been modelled as shown in Figure 8.
This indicates that if the ventilation rate is increased then the AWI gets
-20
-10
0
10
20
30
40
50
60
70
80
90
100
Pre 1900 1900 - 1965 1966 - 1976 1977 - 1981 1982 - 1990 Post 1990
Index
AWI
Fuel Rich MI
Fuel Poor MI
Linear (AW I)
Linear (Fuel Rich
MI)
Linear (Fuel Poor
MI)
worse. As the ventilation rate is increased the general trend of the MI is
positive. However, in dwellings occupied by the fuel poor the MI gets better
until the air change rate reaches 1 air change per hour. Beyond this point the
risk of mould growth increases significantly. This is shown in Figure 8
below.
Figure 8 - Effect of ventilation rate on AWI and MI
References
Department of the Environment "English Housing Condition Survey 1991:
Energy Report" (HMSO) (1996)
International Energy Agency (IEA) "Annex 14: Condensation and Energy:
Volume 1: Sourcebook" (IEA) (1991)
Adan O. C. G, "On the fungal defacement of interior finishes", PhD thesis,
Netherlands Organisation for Applied Scientific Research (TNO), (1994)
Arlian L, "Water balance and humidity requirements of house dust mites",
Exp. & Applied Acarology, 16, 15-35, (1992)
Ashmore I, "Asthma, housing and environmental health", Environ Health,
January 1998, 17-23
Bordass W & Oreszczyn T, "Internal Environments in Historic Buildings:
Monitoring, Diagnosis and Modelling", Report to English Heritage, October
(1998)
Boyd D, Cooper P, Oreszczyn T, "Condensation risk prediction: Addition of
a condensation model to BREDEM", Building Serv Eng Res & Technol,
9(3), 117-125, (1988)
BS5250, British Standard Code of practice for control of condensation in
buildings, British Standard Institute, (1989)
Bronswijk J, "House Dust Biology", Zoelmond, Netherlands, (1981)
-20
-10
0
10
20
30
40
50
60
70
80
90
100
0.4 0.6 0.8 11.2 1.4 1.6 1.8
Ventilation Rate (ach)
Inde x
Affordable Warmth Index
Fuel Rich MI
Fuel Poor MI
Linear (Affordable Warmth
Index)
Poly. (Fuel Rich MI)
Poly. (Fuel Poor MI)
Cunningham M J, "Controlling Dust Mites Psychrometrically - A Review for
Building Scientists", Indoor Air, 249-258, (1996)
Humfrey C et al, "Indoor Air Quality in the Home", MRC Institute for
Environmental Health Assessment A2, (1996)
Korsgaard J, "House-dust mites and absolute indoor humidity", Allergy, 38,
86-96, (1983)
Oreszczyn T, “Insulating the existing housing stock: mould growth and cold
bridging”, in Energy Efficient Building (Eds: Roaf S & Hancock M),
Blackwell Scientific Press, (1992)
Oreszczyn T and Pretlove S, "Condensation Targeter II", Building Serv Eng
Res & Technol, 20 (4) (1999)
... It is based on monthly steady-state solutions. Another version of the tool measures the risk of mould on the coldest surfaces within the dwelling each month of the year [139]. WUFI [132] MRD The results from WUFI are used as input for the calculation of onset of mould growth according to MRD. ...
Article
This paper presents a systematic literature review about the development of criteria and models representing mould growth in wood-based materials. First, results from experimental research regarding factors governing mould growth are discussed; afterwards, they are used to analyse the comprehensiveness of current mould models. The review shows substantial discrepancies between criteria reported for mould growth. Moreover, mould models differ with respect to governing factors and their interrelations, applied methodology, experimental set-ups and nutrients, and how they express mould. Lastly, this paper proposes solutions that account for or reduce uncertainties related to the representation and design against mould occurrence.
... Studies have highlighted that at present many people live in conditions that are considered 'unhealthy' because of high levels of mould growth, damp and dust mite concentrations. Approximately 20% of all dwellings in England suffer from mould growth and dampness to some degree (Oreszczyn and Pretlove 2000). Oreszczyn and Pretlove point out that a key parameter in house dust mite survival and mould growth is relative humidity. ...
Book
Full-text available
Integrated approaches are required to achieve built environment sustainability. However, there is little evidence of research into integration for sustainability, in particular the technology and process of integration in a multidisciplinary setting. Questions may be asked as to the need for integration: is it a response to the inadequacies of stand-alone approaches or merely a diversion from the poorly implemented single sector policies. Sustainability has traditionally been looked as a way of satisfying certain indicators linked with performance criteria; be it environmental, economic or social. Such 'criteria satisfaction' efforts usually overlook the process of achieving sustainability, hence affect the success of a policy or project. This international conference is aimed at bringing together sustainability stakeholders from a wide variety of disciplines and sectors into a platform for discussions on integrated approaches for sustainability. 24 contributed papers and one keynote paper that discuss various approaches for enhancing sustainability in the built environment are presented in this collection. Published papers represent current trends in research into integrated approaches for sustainability.
Article
Energy used in dwellings is an important target for actions to avert climate change. Properly designed and implemented, such actions could have major co-benefits for public health. To investigate, we examined the effect of hypothetical strategies to improve energy efficiency in UK housing stock and to introduce 150 million low-emission household cookstoves in India. Methods similar to those of WHO's Comparative Risk Assessment exercise were applied to assess the effect on health that changes in the indoor environment could have. For UK housing, the magnitude and even direction of the changes in health depended on details of the intervention, but interventions were generally beneficial for health. For a strategy of combined fabric, ventilation, fuel switching, and behavioural changes, we estimated 850 fewer disability-adjusted life-years (DALYs), and a saving of 0.6 megatonnes of carbon dioxide (CO(2)), per million population in 1 year (on the basis of calculations comparing the health of the 2010 population with and without the specified outcome measures). The cookstove programme in India showed substantial benefits for acute lower respiratory infection in children, chronic obstructive pulmonary disease, and ischaemic heart disease. Calculated on a similar basis to the UK case study, the avoided burden of these outcomes was estimated to be 12 500 fewer DALYs and a saving of 0.1-0.2 megatonnes CO(2)-equivalent per million population in 1 year, mostly in short-lived greenhouse pollutants. Household energy interventions have potential for important co-benefits in pursuit of health and climate goals.
Article
Full-text available
Recent studies show that the UK is one of the countries with the highest asthma prevalence worldwide. It has been suggested that the trend towards the reduction of domestic ventilation rates for energy-efficiency reasons has resulted in poor indoor air quality and may represent a causal factor for the high asthma prevalence in the UK. This study firstly aims to assess whether a link exists between asthma and low ventilation rates in housing. Secondly, the study aims to establish the minimum ventilation rate required in a dwelling in order to control levels of moisture-related pollutants (dust mites, mould) and therefore reduce the number of respiratory hazards. The work was funded by the UK Government's Building Regulations Research Programme. An extensive review was performed of the UK and overseas published information on the links between asthma and domestic ventilation rates. The study also included an analysis of existing UK data sets where respiratory health, mould growth and housing characteristics were surveyed. In addition, the study involved theoretical modeling of a typical UK dwelling, to predict the impact that changes in ventilation rates will have on house dust mite (HDM) in a bed and on mould growth. The literature review has highlighted that most existing data is inadequate for conclusions to be drawn regarding the direct association between ventilation rates and respiratory problems. For moisture-related pollutants the literature offers some advice on the minimum ventilation rates required to reduce pollutants levels. The analysis of existing UK data sets has not revealed any significant direct links between ventilation, respiratory health and moisture-related pollutants. The results of the theoretical modeling indicate that in most cases 0.5 ach-1 is required to avoid mould growth but a significantly higher ventilation rate (0.8 ach-1) may be required to control mites’ growth. Most current guidance on domestic ventilation is based on the assumption that if adequate ventilation is provided for avoiding mould growth, then other IAQ problems will be contained. This study indicates that this may not be the case. In order to fully evaluate the direct link between asthma and low ventilation rates in the UK, a large scale prospective study is needed, where indoor pollutants, respiratory health indicators, air-tightness and ventilation rates are all adequately measured.
Article
Since the last decades of the 19th century, technological advances have brought substantial improvements in the efficiency with which energy can be exploited to service human needs. That trend has been accompanied by an equally notable increase in energy consumption, which strongly correlates with socioeconomic development. Nonetheless, feasible gains in the efficiency and technology of energy use in towns and cities and in homes have the potential to contribute to the mitigation of greenhouse-gas emissions, and to improve health, for example, through protection against temperature-related morbidity and mortality, and the alleviation of fuel poverty. A shift towards renewable energy production would also put increasing focus on cleaner energy carriers, especially electricity, but possibly also hydrogen, which would have benefits to urban air quality. In low-income countries, a vital priority remains the dissemination of affordable technology to alleviate the burdens of indoor air pollution and other health effects in individuals obliged to rely on biomass fuels for cooking and heating, as well as the improvement in access to electricity, which would have many benefits to health and wellbeing.
Thesis
Full-text available
In the present work, fungal growth is studied both under steady-state and tran¬sient indoor conditions. The frequent occurrence of fun¬gal problems in Dutch indoor environments that are considered relatively dry, probably emphasises the significance of dynamics, the conse¬quences of mois¬ture storage in particular. Considering this every¬day reality, the main objective of the research described in this thesis was to im¬prove the understanding of the pro¬cess inducing the fungal defacement of inte¬rior finishes. Such improved insight should be a first step to¬wards eco-friendlier, healthier and lasting control strate¬gies of indoor fungal growth. In order to prevent to many parameters ob-scuring the basic process, the experimental work in this thesis concerns a single fungal species only, i.e. Penicillium chrysogenum (CBS 401.92), and is limited to gypsum-based materials. The thesis starts with an introduction into the abiotic factors affecting fungal growth, outlining the conditions occurring in building practice. Although fungal disfigurement of materials is often related to surface condensation, experimental evidence shows that most fungi readily germinate and grow on substrates in equilibrium with relative humidities (RH's) below saturation. Moreover, virtually all fungi have optimum conditions for growth in this range. Chapter 2 is particularly concerned with the vital role of water, defines the quan¬tities commonly used to describe water relations and discusses their background within the context of the water uptake mechanisms of the fungus. The widely accepted osmotic concept, underlying the water relations of cells, is compared to a radically different theory based on multilayer polarization. In contrast to the osmotic model that essentially assumes water uptake in the liquid phase, this new theory applies to the uptake of water in the vapour phase. The Chapter concludes with a preliminary investigation into the response of the fungal cell to transient water vapour pressures using Environmental Scanning Electron Microscopy. This new technique allows real-time SEM-observation of the fungal cell response. The results indicate that conidia of P. chrysogenum are capable of an instantaneous water vapour uptake as the RH increases, suggesting that short periods of high RH should not be neglected in evaluation of the indoor climate with respect to fungal growth. Chapter 3 consists of three parts, dealing with fungal growth under steady-state conditions. In Part I, Standards for testing fungal resistance of materials are reviewed and a new method is proposed to test interior finishes. In Part II, a modified cryogenic method is introduced to observe fungal growth in its most natural form on the substrates considered. In spite of the high porosity of the gypsum, observations suggest that at least the initial growth of P. chrysogenum should be considered a superficial phenomenon. In Part III, the effect of material properties on fungal growth is investigated sys¬tematically on the basis of the proposed testing method. The research in Part 3 contrasts with the usual fungal growth experiments, applying to an overall as-sessment of the fungal resistance and not focusing on the resistance and material properties relationship. It is con¬cluded that the actual fungal resistance of the decorative finish in practice may also depend on the characteristsics of the base substrate. Moreover, the results indicate that the fungal resistance of the multilayered system may be considerably lower than the fungal resistance of the constituent materials.The LTSEM obser¬vations in Part II are consistent with these results. The effects of indoor humidity dynamics on fungal growth are dealt with in Chapter 4 and 5. Chapter 4 concerns the water balance of the gypsum substrate, and par¬ticularly focuses on the moisture storage during wet periods causing prolongation of surface humidi¬ties favourable to growth during dry periods. It is generally recognised that the water trans¬port in monolithic porous materials can be de¬scribed using a macroscopic diffu¬sion-type equation. Until now, the de¬ter¬mination of the diffusion coefficients in this equa¬tion was a major problem. Two new non-destructive measuring tech¬niques were developed on the basis of the neutron beam attenuation and nuclear magnetic resonance principle, respec¬tively. Both techniques are applied to de¬termine the isothermal moisture diffu¬sivity directly from transient moisture con¬tent profiles. For the first time, such profiles in a dry material responding to transient relative humidities below satu¬ration are measured. The results show that the moisture diffusivity to be used in the modelling of moisture transfer is strongly dependent on the initial conditions of the material, underlining the complex in¬teraction of liquid and vapour phases in the porous system. The moisture diffu¬sivity for drying of the wet gypsum may ex¬ceed the moisture diffu¬sivity of the dry gypsum exposed to transient RH's by a factor of 100. As a con¬sequence of the low value of the latter, no hygroscopic inertia ef¬fects are ob¬served, implying that the surface RH hardly differs from the transient RH in the adjacent air. In this case, the experimental results refute the common estimations of hygroscopicity effects on the basis of the apparent diffu¬sivity de¬rived from the sorption isotherm and water vapour permeability. Taking account of these results, experiments on fungal growth responding to transient RH's below saturation are performed (Chapter 5). Despite the everyday reality of climate dynamics, no data with respect to this issue have been found in the literature. A new experimental method to study these effects is elaborated, including non-linear regression techniques to model sigmoidal curves describing vegetative fungal growth. The statistical analysis of transient RH effects on fungal growth is based on response variables deduced from these curves. In this Chapter, the time-of-wetness (TOW) is introduced as an overall measure for the water availability to fungal growth under transient conditions. The TOW is defined by the ratio of the cyclic wet period, i.e. when the RH≥80%, and the cyclic period. The preliminary experiments indicate that growth of P. chryso¬genum on the gypsum-based finsihes is only weakly affected for a TOW≤0.5, whereas it accelerates strongly with increasing values>0.5. The RH value during the drying periods shows hardly any influences on the fungal growth-TOW rela¬tion. Furthermore, except for very fast oscillations of the RH, the frequency of high RH periods only slightly affects the TOW effects on fungal growth. Finally, Chapter 6 comments on the effects of the thermal quality of the building construction, the average indoor air humidity and airborne water vapour trans¬port on the risks of fungal defacement of interior finishes. A temperature ratio criterion is considered a universal step towards reduced risks of fungal deface¬ment of interior finishes at the building envelope, but reduction of the TOW is an effective and straight-forward tool appropriate in the case of climate dynamics. When the dampness of gypsum finishes is related to the indoor air, prolongation of the TOW only occurs in the case of surface condensation. In conclusion, a perspec¬tive of envisaged control strategies is given. Basic princi¬ples for tackling the problems of fungal growth are related to an integral approach of the indoor environment, the building envelope and the interior finish. It may be apparent that the TOW concept supports two main areas of development for reducing the risk of fungal defacement. Besides the obvious improvement of ven¬tilation efficiency, the potentials of material modification to shorten the TOW should be considered. The issue of the fungal resistance of interior finishes should be part of future development towards evaluation tools of indoor fungal growth.
Article
The paper describes a design tool for assessing condensation risk in dwellings and the effect of remedial measures thereon. The BREDEM energy model is augmented by a moisture model to determine mean internal relative humidity (MIRH). This measure of condensation risk is calculated for two zones in a dwelling from mean internal temperatures, moisture generation and ventilation rates. Primary input data relate to occupancy (fuel expenditure and moisture production) and dwelling characteristics (thermal and ventilation). MIRH results are presented as a function of space heating input in example dwellings with remedial measures applied (insulation, draught stripping and extract fans). The constraints imposed by household income and the implications for condensation risk are discussed.
Article
Abstract The literature for the control of dust mites by modification of the psychrometric conditions of the environment is reviewed from the standpoint of a building scientist or engineer, both to present to building science workers an envelope of micro-environment psychrometric conditions to use as control, and to highlight those areas of dust mite biology that require further research to complete the knowledge of the psychrometric envelope for dust mite viability. Some important data to allow tight specification of psychrometric control conditions are missing, viz.: the temperature dependence of critical equilibrium activity for Dermatophagoides pteronyssinus and Euroglyphus maynei; behaviour of dust mite populations under a fluctuating climate; and the difference between wild and laboratory populations. The widely quoted figure for dust mite control of 7 g/mg absolute humidity should be used with caution.
Article
The house dust mites, Dermatophagoides farinae, D. pteronyssinus and Euroglyphus maynei, are prevalent in homes in humid geographical areas throughout the world. These mites thrive in humid environments in human dwellings where there is no liquid water to drink. However, their bodies contain 70-75% water by weight, which must be maintained in order to reproduce. Their primary source of water is water vapor which is actively extracted from unsaturated air. At relative humidities above 65-70%, adequate amounts of water can be extracted from unsaturated air to compensate for that lost by all avenues. Active uptake is associated with ingestion of a hyperosmotic solution which is secreted by the supracoxal glands. Active mites do not survive longer than 6-11 days at RHs < or = 50%. They survive extended dry periods by forming a desiccation-resistant protonymphal stage which can survive for months at RHs below the critical humidity for active stages. Feeding rate and allergen production is directly influenced by RH. Mites feed, multiply, and produce more fecal matter at higher RHs than at lower ones.
Article
The concentration of house-dust mites (Dermatophagoides spp.) was investigated for four seasons in three locations in each of 50 Danish apartments. Simultaneously the absolute humidity was recorded and the previously known correlation between mite counts and indoor humidity was confirmed. It appeared, however, that apartments which had a low absolute indoor humidity in the winter period (due to low household load of water vapour) did not contain noticeable concentrations of house-dust mites in the summer and autumn despite the fact that the indoor absolute humidity in these apartments could be high enough to allow for a high peak-population of mites. Because of this it is suggested that in a temperate climate avoidance measures against house-dust mites should be supplemented at least by a drying out period in the winter, when this process is convenient to perform because of the low outdoor absolute humidity.
Asthma, housing and environmental health
  • I Ashmore
Ashmore I, "Asthma, housing and environmental health", Environ Health, January 1998, 17-23
Internal Environments in Historic Buildings: Monitoring, Diagnosis and Modelling
  • W Bordass
  • T Oreszczyn
Bordass W & Oreszczyn T, "Internal Environments in Historic Buildings: Monitoring, Diagnosis and Modelling", Report to English Heritage, October (1998)
Condensation risk prediction: Addition of a condensation model to BREDEM BS5250, British Standard Code of practice for control of condensation in buildings
  • D Boyd
  • P Cooper
  • T Oreszczyn
Boyd D, Cooper P, Oreszczyn T, "Condensation risk prediction: Addition of a condensation model to BREDEM", Building Serv Eng Res & Technol, 9(3), 117-125, (1988) BS5250, British Standard Code of practice for control of condensation in buildings, British Standard Institute, (1989) Bronswijk J, "House Dust Biology", Zoelmond, Netherlands, (1981)
Insulating the existing housing stock: mould growth and cold bridging
  • T Oreszczyn
Oreszczyn T, "Insulating the existing housing stock: mould growth and cold bridging", in Energy Efficient Building (Eds: Roaf S & Hancock M), Blackwell Scientific Press, (1992)