Sunteți pe pagina 1din 38

Textile Progress

ISSN: 0040-5167 (Print) 1754-2278 (Online) Journal homepage: http://www.tandfonline.com/loi/ttpr20

Clothing systems for outdoor activities


Matthew P. Morrissey & Ren M. Rossi
To cite this article: Matthew P. Morrissey & Ren M. Rossi (2013) Clothing systems for outdoor
activities, Textile Progress, 45:2-3, 145-181, DOI: 10.1080/00405167.2013.845540
To link to this article: http://dx.doi.org/10.1080/00405167.2013.845540

Published online: 19 Feb 2014.

Submit your article to this journal

Article views: 733

View related articles

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=ttpr20
Download by: [Politecnico di Torino]

Date: 27 September 2016, At: 10:39

TTPR_A_845540.3d (TTPR)

14-02-2014 17:21

Textile Progress
Vol. 45, Nos. 2-3, 2013, 145181

Clothing systems for outdoor activities


Matthew P. Morrissey* and Rene M. Rossi
Empa Swiss Federal Laboratories for Materials Science and Technology, Laboratory for
Protection and Physiology, Lerchenfeldstrasse 5, CH-9014 St. Gall, Switzerland
(Received 4 February 2013; final version received 25 July 2013)
Participation in outdoor activities can improve mental, physical and social well-being.
Such activities also present significant physiological strain and risks such as hypothermia; therefore, correct choice and usage of clothing is extremely important. The aim
of this review is to critically analyse the literature regarding outdoor clothing systems,
focusing on the layers comprising a typical clothing system. Additionally, alternative
systems, potential improvements and future trends are discussed.
Keywords: Outdoor activities; performance clothing systems; human safety and
comfort; product design

1. Introduction
The number of participants in recreational outdoor pursuits such as land-, winter- and
water-based activities is large and growing [1]. The growing popularity of participation in
outdoor recreation activities is perhaps due to the numerous psychological and physical
benefits associated with such activities. Outdoor recreation is a major factor in human
wellness, a combination of physical, mental and social well-being. In green environments, people perceive better general health, and spending time in such places has been
correlated with lower stress levels and higher amounts of physical activity [2]. Low stress
and physical activity have been shown to reduce many common health problems such as
high blood pressure, obesity, heart attacks, cancer and mental health problems [3]. Spending time in natural places is increasingly being recognised as an important preventive
medicine with positive health, mental, social and environmental outcomes [4,5].
However, activities such as hill walking can present a significant metabolic and thermoregulatory strain, as well as risks such as hypothermia: though with the proper management of appropriately selected clothing, these risks can be minimised [6,7]. With the
correct clothing the benefits of outdoor pursuits can be enjoyed regardless of adverse
weather conditions, leading to the typical Scandinavian adage there is no such thing as
bad weather, only bad clothing.
The benefits of improving protective clothing for outdoor recreation therefore include
encouraging enjoyable and safe participation at a grassroots level, and perhaps also
increasing performance and extending the limit of human endeavour in harsh environments. Innovations and improvements in this area may also have implications in other
areas such as natural disaster management (e.g. avoiding heat stress for emergency workers) military applications, emergency and rescue services, and so on.
*Corresponding author. Email: matthew.morrissey@empa.ch
ISSN 0040-5167 print / ISSN 1754-2278 online
2014 The Textile Institute
http://dx.doi.org/10.1080/00405167.2013.845540
http://www.tandfonline.com

TTPR_A_845540.3d (TTPR)

146

14-02-2014 17:21

M.P. Morrissey and R.M. Rossi

Consumers are subjected to a lot of information regarding outdoor clothing but this
is often skewed by industry-bias and marketing-hype [8]. Understanding the science
behind the performance of outdoor clothing may help consumers and manufacturers
make more effective and even more ecological choices regarding clothing for outdoor
recreation, protecting the natural environment in which they choose to spend their
leisure time.
1.1. Human thermoregulation and the need for protection
Humans aim to maintain thermal homeostasis with a deep body temperature of 37 C [9].
Since cell functions are highly dependent on enzymes to catalyse metabolic reactions, the
human body can only operate within quite narrow margins or deviations of 5 C and
10 C from the ideal deep body temperature. Metabolic processes in the human body all
result in heat production: this metabolic heat is a product of the basal metabolic rate
(minimum requirement to maintain vital body functions), muscular activity and the thermal effect of food (digestion) [9]. Since the human body operates in such a relatively narrow temperature range, heat transfer to and from the environment is of great importance.
Heat transfer can occur by [10]:
 Conduction. Conductive heat transfer occurs in a solid object or stationary fluid
(such as trapped air); energy is transferred from more energetic to less energetic
particles by interactions between the particles. Conduction between two objects
relies on physical contact between these objects: in the case of clothing, this may
be between fabric layers in contact with each other or between the fabric and the
human skin.
 Convection. Convective heat transfer occurs between a fluid in motion and a bounding surface with different temperatures; the transfer of energy is due to the bulk or
macroscopic motion, but also may be due to random molecular motion in the fluid.
According to the nature of the flow, convective heat transfer can be classified as
forced where the flow is caused by external means, or free, where the flow is
caused by buoyancy forces due to temperature-driven density differences in the
fluid.
 Radiation. Thermal radiation is emitted by all matter with a non-zero temperature.
Thermal radiation is transferred in the form of electromagnetic waves and does not
rely on a transfer medium as do conduction and convection, i.e. it can occur in a
vacuum and in fact is most efficient in this scenario.
 Evaporation/condensation. Heat transfer may also be due to phase changes of
water, e.g. evaporation and subsequent condensation of sweat.
The human body has various methods of thermoregulation; it exploits the various
modes of heat transfer in an attempt to maintain a stable body temperature. To conserve
heat, the vascular system constricts, cooling the skin temperature (reducing the temperature difference between the skin and the environment), maximising the insulation provided by body fat and changing the thermal conductivity of the dermis by a factor of
410. Muscular activity in the form of shivering produces heat, and during prolonged
exposure to cold, hormones increase the basal metabolic rate. Vascular adjustments for a
nude person with normal body fat provide effective thermoregulation between 25 C and
29 C. Intense muscular activity can sustain a deep body temperature within tolerable limits at temperatures as low as 30 C [11]. To dissipate heat, the vascular system dilates;
with high levels of heat loss 15%25% of cardiac output is directed to the skin. At high

TTPR_A_845540.3d (TTPR)

14-02-2014 17:21

Textile Progress

147

metabolic rates sweating is the main heat-loss mechanism of the body. Water also evaporates from the respiratory tract. Water evaporating from the skin and respiratory tract
extracts 580 kcal l1 from the body. Behavioural, in addition to physiological, adaptions
can be used to control heat transfer to and from the body without resorting to clothing,
e.g. by finding a cool shade or water, or by changing the posture or building shelter, or
using fire to preserve heat.
Ideally, the heat production of the body and heat transfer to the environments should be
equal. This concept is referred to as thermal balance and is summarised by the equation:
M  W RES E R C K S;

where M is the metabolic energy production, W is the effective mechanical energy, and
RES, E, R, C and K represent heat transfer by respiration, evaporation, radiation, convection and conduction, respectively. S is the change in body heat storage. Of course, thermophysiological and behavioural approaches to thermoregulation are not always sufficient to
maintain thermal balance and thus clothing is required. Indeed, genetic analysis of human
body lice suggests that the invention of clothing (beyond that fulfilling decorative and
social functions, which may have occurred earlier) may have coincided with the spread of
modern Homo sapiens from the warm climate of Africa, between 50,000 and
100,000 years ago [12].
Clothing systems, in addition to physiological and behavioural processes, also function by modifying the heat transfer between the human body and the environment. One
example of an ancient, 5000-year-old clothing system is that found on a body frozen in
the Austrian Alps [13]. This clothing consisted of an assortment of furs with fastenings to
modify ventilation and a weather-resistant grass outer layer.
1.2. Modern layering systems
The most commonly used clothing system for outdoor activities is still based on a remarkably similar concept to that used by the ancient man found frozen in the ice. It consists of
a base layer worn for next-to-skin comfort, mid layer primarily to provide insulation, and
shell layer to provide protection from wind and precipitation. Shell layers are usually
made from waterproof breathable fabrics (WBFs) [14]. There is a popular alternative to
the traditional layering system known as soft-shell clothing. The aim of a soft-shell system is to provide protection from cold, wind, rain and overheating without the need to
add or remove layers, and originally without using a waterproof breathable membrane
[15]. This is achieved by combining fabric elements from the traditional layering system,
usually knitted fleece or pile fabrics, and a tightly woven shell layer with or without a
membrane, either by lamination or sewed construction. Original soft-shell clothing sacrificed absolute waterproofness for increased breathability and simplicity of construction
by omitting WBFs; these garments were designed to be worn without a base layer. The
soft-shell category of garments evolved, particularly when large manufacturers of WBFs
began making soft-shell garments, such that many modern soft shells incorporate a membrane and/or coating and additionally, many users combine the soft shell with a base layer.
Considering the fibres and fabrics are of extremely similar construction, the soft-shell
category of garments really represents a design approach and attitude toward dressing for
the outdoors, and therefore the information in this review can be applied to soft-shell
garments and clothing systems.

TTPR_A_845540.3d (TTPR)

14-02-2014 17:21

148

M.P. Morrissey and R.M. Rossi

2. The base layer


2.1. Wetting, wicking and thermophysiology
The vast majority of product literature and many peer-reviewed articles (e.g. [16]) state
that the primary role of the base layer is to wick water away from the skin (transplanar
wicking), and also to spread the sweat over a large surface area (in-plane wicking) for better evaporative cooling and faster drying. The science behind wetting and wicking has
been covered in several previous reviews [17,18]. Improvement of the wetting and wicking properties of textile products can be achieved by a number of techniques [19]:
 By changing the fibre cross section to increase the surface area (e.g. [20,21])
 By blending hydrophilic and hydrophobic fibres into yarns (e.g. [22])
 By combining hydrophilic and hydrophobic yarns in fabrics (usually double-layer
knitted) (e.g. [23])
 By changing the surface hydrophilicity of hydrophobic fibres (e.g. chemical or
plasma treatment) (e.g. [24])
Additionally, the dynamic stretching and relaxation of yarn filaments and yarns during
fabric deformation has been shown to increase transverse wicking via a pumping effect
[25]. Oner et al. [26] and Wardiningsih and Troynikov [27] found that the liquid transport
effectiveness or overall moisture management capacity also decreases with increasing
fabric tightness.
However, as Rodwell [28] suggested in 1965, the value of wetting and wicking from a
thermophysiological perspective is still not clear. Whether such wetting and wicking
properties are actually desirable from a thermophysiological perspective is currently
under debate. Burton and Edholm [29] recognised that because the evaporative cooling
efficiency is decreased when water is absorbed by the fabric (as the loci of evaporation
are farther away from the skin), transplanar wicking fabrics may have a detrimental
effect (i.e. exacerbating the risk of heat stress during exercise). Kerslake [30] also identified that water evaporating from the clothing would have a lower cooling efficiency.
More recent work by Havenith et al. [31,32] has quantified the reduction in evaporative
cooling efficiency as simulated sweat is systematically moved away from the skin
through the clothing layers. For example, the evaporative cooling efficiency of wet
underwear under personal protective equipment (PPE) is around 0.65, whereas that of
wet skin under PPE is close to 1. When the outer layer is wet, the evaporative cooling
efficiency is reduced further to between 0.2 and 0.4. Wang et al. [33] studied this problem with tight fitting one-layer sportswear and found that due to this effect, thinner fabrics are more suitable in terms of maintaining a high evaporative cooling effect. This
concept has also appeared in a popular online backpacking magazine in an article entitled
Just say no to wicking, discussing the pros and cons of conventional wicking base
layers and fishnet base layers that allow sweat to evaporate from the skin surface [34].
The concern regarding the evaporative cooling efficiency has also been used in recent
product marketing, with the manufacturer claiming that their base-layer product offers
the correct balance between allowing optimal evaporative cooling of the skin and
removing excessive liquid from the skin, but this statement is not supported by any
empirical evidence [35].
Despite this potential thermophysiological drawback, wicking base layers may
improve comfort in terms of tactile comfort. Since human skin features no sensor for
moisture or skin wetness sensations (but skin wetness is positively correlated with

TTPR_A_845540.3d (TTPR)

14-02-2014 17:21

Textile Progress

149

discomfort), moisture is detected by temperature and tactile sensors; therefore, effective


transplanar wicking, even though it may reduce the evaporative cooling efficiency, may
improve comfort sensation by reducing the strong tactile sensation caused by the sticky
wet fabric clinging to the skin [36]. Similarly, the sensation of water droplets rolling
down the skin surface will also be reduced or eliminated, and changes in temperature
caused by wet skin will be attenuated.
2.2. Drying time and water absorbed
The amount of sweat or water a base layer absorbs and the time it takes to dry is another
important issue. Commonly, advertisements, popular press and even scientific publications proliferate the idea that fabrics constructed from synthetic materials absorb very little sweat or moisture and therefore leave the skin warm and dry, whereas natural
materials absorb water and contribute to heat debt [37]. Closer inspection of the literature
reveals a less simple picture.
In 1951, Fourt et al. [38] made a thorough and systematic investigation into the drying
behaviour of various fabrics, when the first synthetic fabrics were becoming available in
the consumer market. They characterised some key drying behaviours that seem to be
ignored in some of the newer literature. Fourt et al. found that by and large all fabrics
dry at the same rate; however, the drying time depends on the amount of water the fabric
originally holds. This water-holding capacity depends primarily on the fabric thickness.
Forty-five years later Crow and Osczevski [37] also found a strong correlation between
fabric thickness and water pickup, and water pickup and drying time. They also presented
evidence suggesting that since the amount of liquid water picked up depends on the total
capillary volume, open-structure fabrics pick up less water. Yoo and Barker [39] found
that although hydrophilic surface treatments of hydrophobic fibres change the rate of
water absorption, they do not change the total amount of liquid water absorption. This
property is dictated by the structural characteristics of the fabric such as thickness and
porosity. Laing [40] found that whilst mass and thickness generally relate to values
obtained on the dynamic sweating hotplate, the bulk density is linked to the absorption
responses. Prahsarn et al. [41] used a novel method to show that the microclimate drying
time, or the time-dependent dissipation of accumulated moisture, is most influenced by
the pore characteristics of the fabric, rather than fabric thickness: this highlights the
importance of thin knit structures with unobstructed inter-yarn pores for optimum moisture vapour dissipation. Other research suggests that hygroscopicity desirability (for sensorial comfort) is dependent on whether external pressure is applied: these studies
concerned sock comfort but could be applied to other scenarios such as pressure created
by a rucksack [42,43]
The notion that natural fibres absorb more water is understandable, given their higher
regain, i.e. they can absorb more moisture vapour than synthetic fibres (but the regain of
nylon, for example, is not negligible). Crow and Osczevski found no correlation whatsoever between fibre regain and the amount of liquid water a fabric absorbed, and considered it erroneous to relate the two. However, Fourt et al. pointed out that one factor
influencing water uptake and drying time is the method of support during wetting and
also any mechanical treatment applied; this highlights that the method in which the fabric
is wetted is important: in Crow and Oszevskis study, the fabrics were left overnight sandwiched between two wet sponges. Since there is no standard procedure for wetting fabrics, it can be difficult to differentiate between water absorbed in the fibres and water
entrapped between the fibres. This is somewhat different from the situation in which

TTPR_A_845540.3d (TTPR)

14-02-2014 17:21

150

M.P. Morrissey and R.M. Rossi

Table 1. Summary of effects of hygroscopic and non-hygroscopic fibres on thermophysiology and


comfort.
Significant differences in:

Rodwell et al. [28]

Vokac et al. [44]

Holmer [66]

Nielsen and
Endrusick [67]

Bakkevig and
Nielsen [49]
Ha et al. [63]

Ha et al. [69]

Gavhed and
Holmer [68]
Kwon et al. [70]

Gavin et al. [47]


Tanaka et al. [71]

Roberts et al. [48]

Wickwire et al. [64]


Laing et al. [72]

Van den Heuvel


et al. [65]

Garments compared

Skin T

Core T

Subjective
response

Wool, terylene
Higher with wool when moving from
31.1C, 26% -> 4.1C, 93% RH
Cotton, polypropylene
Lower with cotton but insulation
also lower: 0.27 vs. 0.29
Wool, nylon
Higher Skin T with wet wool than
with wet nylon, only at rest
Cotton, wool, polypropylene,
polyester
Only differences in sweat
accumulation were observed
Wool, polypropylene
Wool feels more itchy but no
difference in thermal comfort
Cotton, polyester
Sweat rates higher with polyester,
clothing T higher with cotton
Cotton, polyester
Lower with cotton, attributed to
higher conductivity due to
moisture absorption
Wool, synthetic
0.3 C higher with wool
Wool/cotton, cotton, polyester
Differences observed in the second
exercise period with 1.5 m s1
wind began
Cotton, synthetic
Cotton, polyester
Skin T lower with cotton; core
T lower with polyester
Thin synthetic, thick synthetic,
cotton
Skin T higher with cotton, but cotton
is twice as thick as thin synthetic
Cotton, polyester

Wool, wool/polyester, polyester


Smaller changes in core T with wool,
lower skin T and better subjective
responses but not significant
Merino wool, cotton, polyester

TTPR_A_845540.3d (TTPR)

14-02-2014 17:21

Textile Progress

151

fabrics absorb evaporated and liquid sweat in a clothing system, and may account for contradictory observations made in wearer trials with human subjects.
In some wearer trials, such as those of Vokac et al. [44] (2 h intermittent exercise
at 2 C), no significant differences in sweat absorption were found between various
combinations of cotton and polypropylene two-layer clothing. However, in many wearer
trials comparing garments made from synthetic and natural hygroscopic fibres, the
hygroscopic fibres absorb more sweat. Holmer [45] found that in wearer trials [8 C, 60%
relative humidity (RH), 0.3 m s1] wool garments accumulated an average 245 g of sweat
and nylon garments 198 g. However, it is not clear to what degree this is due to fibre
regain, since Holmer reported that the wool garment had a slightly higher thickness and
thermal resistance. Li et al. [46] found significant differences in sweat absorption
between T-shirts of different fibres (wool, cotton, acrylic, polypropylene, polyester
(PES), acrylic, viscose, poly-cotton) at cold temperatures (14 C, 32% RH) but not in
warm temperatures (32 C, 45% RH); however, as this was not the main interest of this
study, the exact values were not reported. Gavin et al. [47] found that cotton retained significantly more sweat than a synthetic fabric (30.5 g vs. 10.5 g) in warm conditions
(30 C, 35% RH, 311 km h1). The two garments were estimated to have a similar thermal insulation but that of the cotton was slightly higher (0.28 Clo vs. 0.27 Clo); therefore,
the cotton fabric may have had a slightly larger thickness. Roberts et al. [48] also conducted wearer trials (20 C, 47.5% RH, 3 m s1) and found that a hot conditions base
layer absorbed less sweat than a cold conditions base layer and cotton base layer: the
amount of sweat was correlated with the thickness for each fabric, but the differences
were not significant (44 g, 0.65 mm; 52 g, 1.00 mm; and 66 g, 1.20 mm, respectively).
Bakkevig and Nielsen [49] found that more sweat accumulated in wool underwear (39 g)
than in polypropylene underwear (10 g), but the wool underwear was 1.95 mm thick and
the polypropylene underwear 1.41 mm. Interestingly, in contrast to some studies, the
open-structure fishnet underwear absorbed more than the normal polypropylene underwear, but the thickness was more than 1 mm greater (2.45 mm). Data from Anderson
[50] regarding the perceived drying time of different garment types provides some evidence supporting the idea that the open-structure fishnet has quickest drying time: the
cotton garment was perceived to be dry in 110 min, wool in 91 min, PES garments in 55
85 min, polypropylene in 48 min and fishnet polypropylene in 24 min. Anderson did
not measure the initial amount of water absorbed.
Another important point may be that wool feels less wet than synthetic fibres
because more water can be bound to polar sites in the fabric structure, rather than being
present in liquid form: therefore, despite having absorbed more water it may feel the
same as a synthetic fabric that has absorbed less moisture. This has been observed in
experiments; however, in one report this effect was only statistically significant at 25%
RH: such conditions are unlikely in real-use or inside a real multi-layered clothing system [51,52]. In other studies, human subjects have been unable to differentiate between
different types of wet fabrics [53]. In the study of Li et al. [54], 10 male subjects wore
wool or PES long-sleeved T-shirts and acclimatised to 28 C and 30% RH. After resting
they walked at 5.6 km h1. The wool T-shirts absorbed four times more sweat than
PES, and Li et al. argue that because of this the wool was perceived as less clammy
than PES. Finally, Fanguero et al. [55] studied the drying properties of a variety of fabrics incorporating wool, Coolmax and Finecool. Although it is not clear whether results
were statistically significant or not, they found that the fabric ranking of drying time
was different at normal (20 C) and high (33 C) temperatures. This could have implications for which fabrics are suitable for drying during wear (at body temperature) or for

TTPR_A_845540.3d (TTPR)

152

14-02-2014 17:21

M.P. Morrissey and R.M. Rossi

drying after wear (at room temperature). However, more research is required to elucidate this behaviour.
In conclusion, it appears that when a fabric is submerged in water (e.g. in the papers of
Fourt et al. [38] and Crow & Osczevski [37]) the amount of liquid water is mainly dependent on the thickness, or in the case of open-structure fabrics, the capillary volume. The
thickness and capillary volume are most likely strongly correlated in conventional fabrics,
and therefore the correlation of water content should be made with capillary volume
rather than thickness. In wearer trials, though it is hard to separate the confounding effects
of fabric thickness and fibre type, it appears that fabrics made of fibres with higher regains
do absorb more moisture, presumably because they gain weight by absorbing both evaporated and liquid sweat, and therefore will take longer to dry.
2.3. Physiological effects of the heat of sorption
Another issue extending from the water absorption and drying time debate is whether the
absorbed water has a physiological effect, either in terms of increased fabric conductivity
due to absorbed water or in terms of the heat of sorption, particularly in the case of wool,
but other fibres such as regenerated viscose also have high regain and thus the heat of
sorption [56]. Knowledge about the regain of different clothing materials was originally
important for textile traders in order to know the true weight of the product, and for technical aspects of spinning; however, the hygroscopic qualities were also expected to influence human comfort and physiology, mainly through the heat of sorption [57]. This effect
led to the recommendation of wool, a minor producer of heat, over plant or synthetic
fibres for cold weather clothing [58].
Numerical models simulating the effects of hygroscopic and non-hygroscopic clothing also suggest that hygroscopic buffering is of little physiological significance. For
example, Farnworth et al. [59,60] argue that non-hygroscopic clothing may in fact be better in maintaining constant body temperature, as the heat of sorption released when sweating occurs will reduce the amount of heat loss possible from the skin, at the very moment
when heat loss is desired. The model of Fengzhi and Li [61] also predicts higher skin temperatures when wearing hygroscopic clothing.
Gibson [62] observed that the temperature of wool increased by up to 12 C at a constant ambient temperature with the RH changing from 0% to 100%. Rodwell [28] showed
that at a constant temperature of 22.8 C when the RH is increased from 30% to 90%, up
to an 8 C rise in garment temperature is observed; however, when the increase in humidity (43% to 93%) was accompanied by a decrease in temperature (20 C to 6.1 C), the
temperature of the wool garment simply minimises the size of the clothing temperature
drop (by 1.5 C compared to a synthetic terylene garment).
Table 1 shows the main results of studies with human subjects regarding hygroscopic
and non-hygroscopic clothing in chronological order. At first glance, it is clear that neither
hygroscopic clothing nor non-hygroscopic clothing has significant influence on human
thermophysiology.
The studies of Ha et al. [63], Gavin et al. [47], Wickwire et al. [64] and Van den
Heuvel et al. [65] found no significant differences in skin or core temperature or subjective responses. These studies were all conducted in warm conditions: (37 C, 60% RH,
0.1 m s1); (30 C, 35% RH, 311 km h1 wind); (35 C wet bulb temperature); and
(41.2 C, 29.8% RH, 4 km h1 wind), respectively. Participants in these studies rested for
90 min; rested, ran, walked and recovered for a total of 60 min; performed simulated
industrial tasks for 120 min; and walked for 120 min before performing an alternating
runningwalking protocol, respectively. In the latter two studies of Wickwire et al. and

TTPR_A_845540.3d (TTPR)

14-02-2014 17:21

Textile Progress

153

Van den Heuvel et al., the synthetic or natural underwear was worn beneath additional
military wear and body armour.
Other studies by Rodwell et al. [28], Vokac et al [44], Holmer [66], Nielsen and
Endrusick [67], Gavhed and Holmer [68] and Roberts et al. [48] found differences only
in the skin temperature due to the heat of sorption, but no significant changes in core temperature or subjective response. Rodwell et al. found slightly higher skin temperatures
with wool in very specific conditions (sweating slightly at rest in a hot dry environment,
31.1 C, 26% RH, moving into a cold damp environment, 4.1 C, 93% RH). Holmer only
found significantly higher skin temperatures with wet wool compared to wet nylon when
at rest in cool temperatures (8 C, 60% RH, 0.3 m s1), while Gavhed and Holmer found
higher skin temperatures with wool compared to synthetic (0.3 C) during exercise in
cold conditions (10 C, 100% RH, 0.2 m s1). The results of the studies of Roberts et al.
and Vokac et al. are hard to interpret due to differences in fabric thickness. For example,
in neutral conditions (20 C, 47.5% RH, 3 m s1 wind), Roberts et al. observed higher
skin temperatures with cotton underwear than with thin synthetic underwear, but the
thickness of the cotton underwear is almost double that of the synthetic (1.20 mm vs.
0.65 mm). Vokac et al. admit that the systematically lower (1 C) back skin temperature
observed with the cotton garment in their study cannot be definitively attributed to fibre
hygroscopicity as the thickness and thermal resistance of the garment were lower
(2.6 mm vs. 3.2 mm; 0.41 Clo vs. 0.62 Clo).
The studies of Ha et al. [69], Kwon et al. [70], Tanaka et al. [71] and Laing et al. [72]
found significant differences in skin or core temperature or subjective responses with natural and synthetic fibres. With the exception of the study of Tanaka et al., these studies
found that the core and skin temperatures were lower and subjective responses more
favourable when subjects exercised in garments made from natural fibres. Kwon et al.
and Ha et al. attributed these differences to the fact that natural fibres absorb more moisture, and therefore their thermal conductivity increases by a greater amount, as the thermal conductivity of water is nearly 20 times greater than that of most textiles. In the
study of Kwon et al., significant differences were only observed when 1.5 m s1 wind
was introduced, after the garments were wetted with sweat.
In the study of Tanaka et al. [71], the effects of PES and cotton garments on subjects
immersing their legs in water of increasing temperature were investigated: they found
that the skin blood flow, skinclothing microclimate and clothing surface temperature
were higher with cotton than PES and these physiological effects were accompanied by
less comfortable sensations. The authors suggest that because of their method where heat
stress is applied to totally stationary participants, heat of sorption effects are only due to
sweating, and changes in clothing fit and body movement induced air flow do not mask
these effects, making their method superior.
Confounding factors often make it difficult to reach solid conclusions about the different effects of fibre type. For example, Wang et al. [73] compared two four-layer clothing
systems [traditional and moisture management function (MMF) clothing]. Participants
walked on a treadmill at 15 C. Wang et al. found that the MMF clothing resulted in
lower humidity next to the skin and that the sweat produced was more effectively transferred to the environment with hygroscopic clothing. The difference between the MMF
and traditional clothing appears to be that the inner layers of the vest and coat were made
from a woolcotton blend rather than nylon; however, the air permeability of these fabrics
differed by a large degree (<0.02 vs. <78.80 ml s1 cm1 at 100 Pa). Therefore, it is difficult to conclude which of these physical parameters (air permeability or fabric regain)
constitutes moisture management. In the study of Guo et al. [74], two types of fabric
were compared: PPE1 consisted of a cotton scrub suit worn inside a 100% polyethylene

TTPR_A_845540.3d (TTPR)

14-02-2014 17:21

154

M.P. Morrissey and R.M. Rossi

barrierman and PPE2 consisted of a waterproof breathable gown and a scrub suit
treated to allow moisture to be transferred away from the skin to the surface of the garment, where it could evaporate quickly to keep the body dry and comfortable. Whilst the
latter, PPE2, did indeed improve thermophysiological comfort, the design of the two garments was not identical: PPE1 was a relatively tight fitting one-piece suit and PPE2 was a
more loose fitting smock ending just below the knees, which would tend to encourage
heat and mass transfer by ventilation [75]. The respirator design was also different. Such
confounding variables again make it difficult to confirm the efficacy of moisture management underwear. Confounding factors exist not only in terms of fabric parameters,
but also in terms of climate and activity type. The effects of hygroscopicity will be very
different if the body produces water vapour (insensible perspiration) or liquid sweat
(sensible perspiration).
Models and laboratory testing often propose that hygroscopic or non-hygroscopic
fibres will have a significant effect on human comfort [76], but this is not always evident
in wearer trials. Therefore, it appears that factors such as air layers, perception insensitivity of human subjects and physiological variability mean that especially in hot environments, fibre hygroscopicity does not have such a large impact on human
thermophysiology and comfort as might be expected from properties measured in laboratory tests. Barnes and Holcombe [77] discuss the reasons why fibre sorption may be
important in the laboratory but not in real wearing conditions: one reason is that the magnitude of vapour resistance can be made deliberately large to investigate how well a
model describes sorption effects, but in real-life the magnitude is smaller and thus the
sorption effects are not as important. They go on to explain that the effects of air movement and the wetting of clothing are likely to outweigh sorption effects. This is demonstrated in the work of Stuart et al. [78] who first dried the wool garments in their study to
maximise the heat of sorption, which were mittens in direct contact with the skin. This
demonstrates that the hygroscopic qualities of wool can be manipulated to provide comfortable sensations, but this comes at the cost of a considerable effort for the user, which
may be unreasonable to expect in normal use. Furthermore, laboratory testing has also
shown that differences present in one-layer fabric tests may be nullified when the same
fabric is tested as part of a multi-layer system [79]. Farnworth and Dolhan [80] also found
that differences between polypropylene and cotton underwear were too small to have any
influence on thermal comfort. Lotens and Havenith [81] conclude that in the event that
absorbing clothing is perceived as being more comfortable than non-absorbing clothing,
this is most likely due to tactile properties and differences in liquid management than
because of any effect on body heat transfer. For example, wool fibres with small fibre
diameters generally elicit positive comfort sensations [82]. There is ongoing debate about
the role of fibre hygroscopicity with regard to the post-exercise chill. Though Holmer
[66] showed that participants resting in wet wool garments were slightly warmer, other
unpublished data shows that participants resting in wet cotton garments (vs. PES) suffered
from a more pronounced post-exercise chill. Therefore, it may be the case that for wool,
the heat of sorption outweighs the increase in fabric conductivity when wet, but for cotton
this is not the case. Further work is required to clarify the exact behaviour.
2.4. Fabric structure
So far, the main focus of this section has been the fibre type of base-layer garments and
the relative merits of wicking and wetting properties. However, some authors claim that
the knit structure of underwear is of far more importance than material fibre type [83]. In

TTPR_A_845540.3d (TTPR)

14-02-2014 17:21

Textile Progress

155

Figure 1. Brynje fishnet mesh and regular underwear.

particular, the differences in regular underwear and mesh, or fishnet underwear, are discussed (see Figure 1). The advantages of open-structure underwear in drying time and
transverse wicking have already been addressed.
Roberts and Bertram [84] describe the proper use of mesh underwear: during exercise, outer garments are loosened and unfastened allowing air circulation in the open network of the vest. After exercise the outer layers are fastened and sealed with a sweat rag
or handkerchief, which produces a stable, insulating layer of air next to the skin. In especially cold weather, a second open-structure Brynje garment was used between other
clothing layers. Fonseca [85] showed with a stationary manikin and low external air flow
(e.g. similar to a resting person) that in terms of thermal insulation, mesh underwear
behaves very similarly to conventional underwear.
Nielsen & Endrusick [83] tested five different types of underwear, including fleece,
one-by-one rib-knit and fishnet, and found evidence to support the notion of air circulation
in mesh underwear improving thermal comfort. They found that the physiological
responses in human wearer trials were not correlated with laboratory measures of thermal
and water vapour resistance. Instead, they found that the underwear with most open structure, namely fishnet, resulted in the lowest skin temperature and skin wetness when worn
as part of a three-layer clothing system during 40 min cycle exercise at 5 C (dew point
temperature 3.5 C). They attribute these differences to the open structure that allows air
to sweep directly over the skin. Goldman [86] reports that in physiological trials in 1944,
the string vest had a higher level of acceptability in terms of subjective thermal comfort
perceptions. Ueda et al. [87] studied underwear garments including a mesh base layer and
in some cases recorded lower temperatures at the chest with the mesh base layer.
Open-structure fabrics may also represent an advantage once the underwear has
become wet. In the study of Bakkevig and Nielsen [88], 10 men rested in wet underwear
for 1 h in cold conditions (10 C, 85% RH, 0.1 m s1). Conclusions about the influence of
fibre type were confounded by large differences in fabric thickness, but the authors concluded that the thickness of the underwear had more influence than fibre type, and that a
fishnet structure might have an advantage as there is a less area in contact with the skin:
this was evidenced by higher skin temperatures and more positive subjective responses.

TTPR_A_845540.3d (TTPR)

156

14-02-2014 17:21

M.P. Morrissey and R.M. Rossi

Critchley also recommends Brynje for heavy exercise in cold climates [89], and
reports further benefits of Brynje in hot weather: specifically, by preventing a sweat-saturated shirt from clinging to the skin, the occurrence of mosquito bites can be reduced.
Woodcock [90] also notes that in cold weather, the air gap adjacent to the skin created by
Brynje decreases heat loss from the body when exterior clothing is wet from sweat or
rain. Such air gaps decrease the cooling effect of wet fabrics on the skin [91].
2.5. Transient effects
Kerslake [92] suggested that underwear has a further function, which is to reduce skin
cooling caused by momentary contact with cold outer fabric layers. Even very thin underwear can create the impression of keeping the skin warm, because cold receptors in the
skin are especially sensitive to transient cooling events. Hes et al. [93] suggest that thin
plain knit underwear reduces the thermal impact of a cool outer fabric by about 43% and
thicker rib-knits or PP knits with higher thermal resistance may reduce the effect of thermal absorptivity of the outerwear by up to 11% of its original value.
2.6. Conclusions and future trends
The literature suggests that the base-layer performance is far from simple and depends on
a complex interaction of physiological and environmental factors. It appears that the conclusions made 60 years ago by Andreen et al. [94] that comfort largely depends on fabric
geometry and construction and the manner in which the fabric is worn on the body (i.e.
garment design, see Nielsen et al. [95]) still stand. That is to say that macroscopic differences are more important than microscopic differences (e.g. fibre type), and differences
in thermophysiological response to different fibre materials are either non-existent or simply difficult to detect.
Essentially, the desirability of factors such as wetting and wicking ability, drying time
and hygroscopicity depends on the application. For example, good wetting and wicking
(transverse and planar), especially with thick underwear, may reduce the efficiency of
evaporative cooling (sweating): for a resting person with wet skin looking to conserve
heat, this will be a desirable property. However, for an exercising person that may potentially overheat, it would be more advantageous for the sweat to evaporate from the skin.
Given these conflicting requirements, the optimal situation may be to have a fabric with
switchable wicking characteristics: maximising sweating efficiency when required and
moving water away from the body by wicking when it is not required. For example, surfaces that switch from hydrophilic to hydrophobic with temperature could allow the wicking of water away from the cold skin, but allow sweat to remain on the surface of the hot
skin [9698]. Alternatively, it has been suggested that rather than one type of base layer
being optimal for all conditions, outdoor activists should consider using different types of
base layer for different activities [99].
From a water absorption and drying perspective, the literature suggests that a very
thin, synthetic fabric (low regain) with an open structure should absorb the least water
(when wetted by a mix of liquid and evaporated sweat) and dry most quickly. Of course,
other practical requirements such as protection from sunlight and social modesty must
also be considered.
The role of hygroscopicity in underwear has also been addressed. Conclusions are
hard to draw from the conflicting evidence in the literature, but it appears that depending
on the environmental conditions and exercise type there may be some factors to consider

TTPR_A_845540.3d (TTPR)

14-02-2014 17:21

Textile Progress

157

when choosing natural or synthetic fibres. Particularly for wool fibres, it appears that the
high regain and heat of sorption mean that at rest, when the underwear is wet, there is a
small heat of sorption effect that may result in marginally higher skin temperature, though
these are usually imperceptible to the wearer. When exercising in warm conditions, the
high regain or water absorption of natural fibres appears to reduce the insulation of the
clothing, thus cooling the wearer: these effects are not likely to be significant when worn
as part of a clothing ensemble or in a highly stressful hot environment. There is some
speculation that the heat of sorption caused by sweat absorption increases heat stress on
the wearer: this effect has been observed but only in conditions where the subjects are still
but sweating; in such situations, fibres with high regain and heat of sorption should be
avoided.
In conclusion, choosing the correct base layer requires a careful consideration of the
exact environmental conditions (or changes in these) that are expected to be encountered,
and the type of exercise performed. In the future, advances in technology may create
materials that can intelligently change their wetting, wicking, drying (porosity) and
hygroscopic qualities to most benefit the user in the specific scenario they encounter.

3. The mid layer


The primary role of the mid layer is to provide insulation. Mid layers are commonly
based on synthetic or natural fibre knitted fleece structures. Synthetic and natural nonwoven battings, or down fibres, are also sandwiched between tightly woven materials to
create mid layer garments. The mid layer functions by trapping still air which has a low
thermal conductivity (0.025 W m1 K1). Due to this, most textile materials have a similar specific conductivity (0.042 W m1 K1); the reason this is different from that of still
air is due to heat transfer by radiation, and with dense fabrics fibre conductivity may also
play a role [100]. Due to this simple relationship between clothing insulation and the
thickness of trapped still air, it is often stated that with regard to dry heat transfer, clothing
can consist of any material that can create still air layers. For example, Holmer [101]
states that choice of textiles is more or less arbitrary as long as good insulation is
provided. Goldman [102] states that if the measured value (of thermal insulation) is not
1.57 Clo cm1, repeat the thickness and Clo measurement, since either one or the other
measurement is wrong.
There are of course exceptions to these statements. In 1948, Backer [103] observed
that because textiles are low-density structures, containing a large proportion of air, the
thermal resistance of textiles is primarily dependent on the behaviour of the entrapped air
layer, i.e. although the density of conventional textiles is such that they resist free and
forced convective flow [104], low-density insulation can exhibit large decreases in thermal resistance with increasing air flow. Conversely, fabrics with high air flow path tortuosity [105] and s large fibre surface area [106] minimise air movement and therefore
convective heat loss. Goldman also states that unusually low-density fabrics will cause
exception to his rule by admitting more heat transfer by radiation. Holmer expands and
adds that with sweating and long outdoor excursions, the low hygroscopicity of syntheticbased mid layers becomes important, whereas moisture absorption in natural fibre mid
layers decreases the thermal resistance of the clothing, and particularly when the metabolic rate decreases after exercise, post-exercise chill presents a threat.
Since the effect of clothing hygroscopicity has been discussed previously, this section
focuses primarily on the effect of air movement on insulation.

TTPR_A_845540.3d (TTPR)

158

14-02-2014 17:21

M.P. Morrissey and R.M. Rossi

3.1. Air movement and insulation reduction


In 1938, Winslow et al. [107] showed that the convection constant for clothed humans,
K, varies with the square root of the air velocity, V (cm s1):
p
K 1:043 V :
In 1946, Rees [108] found a similar relationship between thermal resistance of an air layer
R (in TOGs) and air velocity V (miles h1, at sea level):
R

1
p :
0:42 0:81 V

Fourt and Hollies [109] also found that the insulation of the clothing boundary layer,
R, was dictated by the air velocity V raised to the power 1/2 (equal to the square root):
Boundary insulation R

1
:
0:61 0:19V 1=2

It is also possible to analyse decreases in insulation as linear increases in conductance,


and some authors claim that this approach is advantageous [110]. Givoni and Goldman
[111] saw that the effective wind speed (Veff) responsible for reducing clothing insulation
was a combination of the wind speed (V) and body motion (linked to the metabolic
rate M) and used the following equation to describe these effects where increases in
metabolic rate cause a linear increase in effective wind speed:
Veff V 0:004M  105:
Havenith et al. [112] also observed that changes in activity (similar to M) and posture can
be described linearly, whereas changes due to wind are described by a square root function. More recently, empirical models combining the effects of walking, wind speed and
air permeability of the outer layer of clothing have been developed [113]. First, equations
incorporated the effects of walking and wind speed [114]:
Insulation reduction% 0:858e0:05V 0:161w ;
where V is the wind speed and w is the walking speed, in m s1. Such equations are also
available for nude and dressed subjects [115]. Other studies have shown that the reduction
in insulation due to wind is related to fabric air permeability [116]. Correction equations
for insulation, It,r, dependent on body movement, wind and clothing air permeability have
been published by Nilsson et al. and Holmer et al. [117,118] and subsequently expanded
by Havenith and Nilsson [101,119] with the form (see also Figure 2)
h
i
2
0:1434
It;r e0:0512V 0:40:000794V 0:4 0:0639wp
 It;static ;
where V is the air velocity, w is the walking speed and p is the air permeability in
l m2 s1. Depending on the value of It,static, different correction equations should be used
(for insulation from 0 to 0.6 Clo, 1.22 to 184 Clo and 1.49 to 3.46 Clo): more details can
be found in Havenith and Nilssons publication [119].

TTPR_A_845540.3d (TTPR)

14-02-2014 17:21

Textile Progress

159

Figure 2. Nilsson and Haveniths relationship for thermal insulation, body movement, wind and air
permeability.

The combined effect of relatively modest walking (0.77 m s1) and wind speed
(1.0 m s1) has been shown to have heat exchange capacities of up to 161 W m2 by
Bouskill et al. [120]. Studies comparing the magnitude of the bellows effect with moving
thermal manikins and human subjects have concluded that due to the greater complexity
of human movement, ventilation and therefore heat transfer is greater with human
subjects [121].
3.2. Separating the effects of air movement on clothing insulation
Body movement (and potentially body heat, in the case of natural convection) and external air movement reduce the total clothing insulation in various separate ways. These are
summarised in Figure 3 and discussed in more detail in the following section.
3.2.1. Reduction of boundary layer insulation
Air layers form on the surface of textile layers by adhesion of the gas molecules to the
surface; further layers are formed by a second layer of molecules adhering to the first,
and so on [122]. In still air this boundary layer may be up to 12 mm thick but with vigorous motion, reduced to 1 mm. Backer [103] suggested that the reduction of the boundary
layer insulation of textiles is the dominant mechanism of clothing insulation reduction up
to air flows of 2.7 m s1. Yankelevich [123] made a theoretical appraisal of the effects of
air flow on clothing. He stated that in the case where the outer layer is totally impermeable, reductions in thermal resistance are due only to the reduction of the boundary layer
(assuming the insulation is not compressed). As well as the external air flow (wind) affecting the boundary layer insulation, body motion can erode the boundary layer. Nielsen
et al. [124] found that the insulation of the boundary layer decreased by 7%26% when
bicycling and by 35%45% when walking. Where body parts are not covered by clothing,
convective heat transfer is higher [125], and the local thermal resistance of such

TTPR_A_845540.3d (TTPR)

160

14-02-2014 17:22

M.P. Morrissey and R.M. Rossi

Figure 3. Illustration of the separated effects of ventilation.

non-covered body parts can be affected by clothing modifications in other areas. For
example, using a hood stabilises the air around the neck and torso, increasing the thermal
resistance in these areas [126]. Regional convective cooling of clothing has also been
studied using thermal imaging techniques [127].

3.2.2. Body motion


In addition to eroding the boundary layer insulation, body movement creates air movement inside the clothing microclimate. Vokac et al. [128] observed the effect of body
movement induced air flow on physiological parameters; they observed that during exercise, the skin temperature, humidity and heat flow reach a new, high steady state, and
upon cessation of exercise an additional sharp increase in skin temperature and humidity
and a decrease in heat flow occur which was taken as evidence of the bellows effect,
by its absence. Fanger [129] states that walking motion creates relative air velocities of
12 m s1. Using a mass-loss technique, Nishi and Gagge [130] also found that when
walking the average air velocity on the surface of the nude human body was between 1
and 1.91 m s1, and that the air speed is highest at the extremities. In 1947, Belding [131]
found that the air movement created by body motion could reduce the thermal resistance
of a clothing system by up to 50%. Nielsen et al. [124] also investigated other activities,
finding that bicycling and walking reduced the intrinsic clothing insulation (excluding
changes in the boundary layer) by 30%50%. Changes in insulation due to body movement are larger with thick fabrics and large air layers [112].

TTPR_A_845540.3d (TTPR)

14-02-2014 17:22

Textile Progress

161

3.2.3. Interaction of wind and body motion


The effect of body motion eroding the outer clothing boundary layer is reduced in windy
conditions [132]. Nielsen et al. [124] also found that once the surface layer of air is disturbed by an external air flow, further disturbances caused by body motion have only a
small insulation-reducing effect.
3.2.4. Free convection
In environmental conditions with still air, free convection, as opposed to forced convection, is dominant [129]. With regard to fabrics and the layers of air present in the clothing
system, Backer [103] suggests that free convection occurs in air layers larger than 1/200
(12.7 mm), Spencer-Smith [133] states that natural convection is negligible in air layers
smaller than around 8 mm, and Farnworth observed the onset of natural convection at
13 mm [134]. Usually the air spaces in textiles are much smaller than this, so natural convection is limited, but Cena and Monteith [135] have estimated that the slow ascent of
free convection in animal furs proceeds at 0.01 m s1. The authors therefore conclude
that this effect is trivial in relation to the velocity of external air flows. Chen et al. [136]
and Fan and Qian [137] have shown that whilst larger garments trap more air and are
therefore warmer, above a threshold natural convection occurs in the air spaces, and thermal and water vapour resistance begins to decrease.
3.2.5. Compression effects
In high winds, clothing systems are squashed or compressed by the air flow, changing
the thickness of the textile and air layers [123]. Backer [103] suggests this starts occurring
at air flows of 2.7 m s1 and above, and is particularly prevalent with open-structure
under layers. In some studies, compression effects have been eliminated with the use of
rigid metal fly sheets [138]. Compression effects also occur with changes in posture such
as sitting, and looser garments show larger changes in insulation because the relative
changes in geometry are larger [125]. Anttonen and Hiltunen [139] observed that compression of clothing is more pronounced with laminated air-impermeable fabrics which
may be compressed up to 45 mm at the windward side; more permeable fabrics are compressed only by 1035 mm. Stevens [15] also observed that with laminated fabrics, warm
air from the microclimate is removed when gusts of wind momentarily compress the
clothing. Other changes in geometry also occur due to wind, in particular inflation at the
rear of the body [139], where inflated air layers of 6070 mm can occur. Due to such
compression effects, Fan [136] concludes that tighter fitting or tailored garments are preferable in keeping the body warm in cold and windy conditions. 3D spacer fabrics have
also been recommended to avoid compression-induced insulation loss [140].
3.2.6. Penetration of fabric layers
When the outer layer of a fabric system is permeable, wind can penetrate and disturb the
inner air layers, reducing the systems thermal resistance: in non-uniform air flows, this
effect causes local differences in thermal resistance [141]. Using both the heated cylinder
and the hot plate methods, Niven [142,143] found that windbreaks (outer layers with low
air permeability) are, quite intuitively, most effective outside the insulation layers, and
are of special importance for open-structure materials (in this case, an open-cell urethane
rubber). Niven also recommends that clothing designed for windy weather should include

TTPR_A_845540.3d (TTPR)

162

14-02-2014 17:22

M.P. Morrissey and R.M. Rossi

an additional exterior windbreak layer. The use of linings to stabilise air in insulating battings has also been suggested [144]. In some models of clothing insulation, windproof
outer layers are given double-weighting [145]. Fonseca and Breckenridge [138,146] proposed a novel solution to the problem of outer-layer wind penetration, based on the fact
that at the time of writing, the manufacture of extremely impermeable outer layers was
problematic. Their solution was to provide an air gap or channel that directed the air flow
circumferentially around the human trunk, rather than penetrating the inner layers of insulation. They also showed that fabrics with different air permeability exhibit different
threshold air flows where convective cooling becomes significant, and demonstrated with
human trials that their concept of a bypass layer could eliminate the requirement for
absolutely air impermeable outer layers. Kind et al. [144] also explored the Fonseca
bypass layer concept, measuring dynamic pressure on a heated cylinder, on the premise
that using bypass layers is desirable as air-permeable outer layers allow increased water
vapour transfer. They suggest using multiple windbreak sheaths and permeable batting
layers to reduce the dynamic pressure penetrating the clothing. In further work, Kind and
Broughton [147] used a plastic mesh to create a bypass layer: such bypass layers could
potentially be created with knitted 3D spacer fabrics. The concept of a threshold permeability where moving air penetrates the fabric was also observed by Lamb and Yoneda
[148] using a rotating heated cylinder; the penetration pressure was found to depend on
the fabric permeability to the power approximately 2/3. Wind penetration can also depend
on the relationship between the fur or fabric geometry and wind direction, e.g. heat loss
through newborn caribou furs is more than doubled when the air flow is parallel to the
hair axis rather than perpendicular to it [145].
3.2.7. Filtration and ventilation
When the clothed human body, often modelled as a cylinder, is exposed to wind, an
uneven pressure builds up on the outer surface (see Figure 4) [123]. If the external
clothing layer or apertures allow air penetration, air will filter through the clothing from
areas of maximum pressure to areas of minimum pressure, thus filtering both in and out
of the clothing system. This effect is also described by Stuart and Denby [149] who predict that air penetration occurs at primarily 30 either side of the windward stagnation
point, and air is sucked from the leeward wake area. Yankelevich states that the degree
of filtering is larger with loosely fitting clothing. It is important to note that in addition
to this mixing of ambient and microclimate air, convective cooling occurs due to internal circulation and movement of air, which would occur even in a perfectly sealed clothing system [125].
Crockford [75] was the first to adopt trace gas techniques (summarised by Lumley
[150]) to assess the effects of fit garment design, fabric permeability, body movement and
wind on ventilation. In accordance with Yankelevich [123], he found that looser fitting
jackets, open apertures and shorter cut jackets (apertures closer to the main trunk of the
human body) all increase ventilation; these conclusions are supported by den Hartogs
work [151]. Increasing the ventilation rate is also possible by increasing fabric air permeability [152] or by adding foam spacers to the clothing (the latter modification can more
than double ventilation rates) [153]. Danielsson [154] also measured increased convective
heat loss coefficients in clothing with larger free air spaces. In terms of ventilation in air
spaces caused by body motion (movement of fabric relative to body), Satsumoto et al.
[155] found that there was an optimal air space of 10 mm for heat transfer, and heat transfer was reduced with air gaps of 30 mm. Birnbaum and Crockford [132] developed the

TTPR_A_845540.3d (TTPR)

14-02-2014 17:22

Textile Progress

163

Figure 4. Air flow characteristics around a human-like cylinder.

technique and measured ventilation rates of 15115 l min1 in various types of foul
weather clothing. They estimated that 300 l min1 would be desirable for hard labour.
Lotens [153] estimated that 450 l min1 is desirable, but 360 l min1 is perhaps maximum
available ventilation, meaning that fabric properties are also important. Tracer gas techniques have also been applied to cylindrical columns where ventilation rates of fabric systems were found to increase linearly with air speed and in a more pronounced fashion
with highly-air-permeable fabrics [156]. Ueda et al. [87] also performed local ventilation
measurements and found ventilation to be highest at the chest, followed by at the back
and arms. It is also possible to estimate the ventilation rate using thermal manikins [157].
Interestingly, the modelling work by Ghaddar and Ghali [158162] assumes that the
body trunk does not swing during walking, and they therefore conclude that trunk ventilation is not as large as limb ventilation. Vokac et al. [128] suggest that there is in fact bellows ventilation caused by the rotation of the trunk and changing size of clothing air gaps,
particularly at the back. This bellows ventilation mechanism may account for
the discrepancy between the theory of Ghaddar and Ghali and the measurements of
Ueda et al.

TTPR_A_845540.3d (TTPR)

14-02-2014 17:22

164

M.P. Morrissey and R.M. Rossi

3.2.8. Openings and apertures


Openings and apertures in the clothing, whether for ergonomic purposes (neck, wrist and
waist openings) or specifically designed to increase the cooling effects of air flow (e.g.
chest, back and pit-zip openings), provide avenues for ventilation and additional heat
loss. In 1965, Fonseca and Woodcock [163] investigated a model clothing system which
featured an opening; in this case, the aim was to simulate a hood. They found that heat
transfer through this opening was higher with a looser fitting fabric and could be limited
with a fur lining or ruff that stabilised the air around the opening. Lotens and Havenith
[153] found that typically vents increased ventilation by 4060 l min1. Vents can be considered to have the same effect, albeit localised, as air-permeable outer fabrics, i.e. they
allow wind to penetrate and disturb the inner layers of the clothing system. Danielsson
[154] used heated thermistors to measure the air speed inside the clothing microclimate
and found that the internal air speed was 30%40% higher with apertures opened in high
winds. Ueda and Havenith [152] also showed that ventilation rates are higher with open
clothing apertures at the wrist, neck, etc. Ruckman [164] studied the effect of openings of
outdoor clothing and found that they could make small reductions in skin temperature.
The largest opening was the most effective. Ruckman found that during rest the jacket
fabric had more importance in maintaining comfort than openings. Satsumoto [165]
found that in some cases one opening allowed more heat transfer than two openings, and
the air velocity was double with one opening than with two. In the same way as body
movement and wind interact such that their combined effect is less than the separate
effects, vents provide smaller improvements when they are close to openings, e.g. at the
waist [153]. Satsumoto [155] showed that heat transfer, and also air velocity [166], is
already largest nearest openings, so further increases in ventilation cause only a small difference. Fanger [129] points out that local convective cooling can be pleasant or unpleasant depending on whether a person initially feels too warm or too cold.

3.3. Air flow effects in laboratory tests


The thermal and water vapour resistance reducing properties of air flow have long been
observed in laboratory-based fabric-testing methods; in 1947, Fourt and Harrist [167]
noticed that the water vapour resistance of an open-structure mosquito net fabric was
extremely low, which they attributed to air penetration. More recently, Gibson [168] identified that when measured with a guarded hot plate, cover fabrics on insulating materials
(especially open foams) dramatically increase the thermal resistance of the insulation,
indicating that air penetration causes large decreases in thermal resistance. Gibson identified that effects of air penetration are dependent on the thickness, pore size and airpassage tortuosity. Dyck et al. [169] also recognised that air penetration of samples in
guarded hot plate thermal resistance testing can cause large differences in results which
can be problematic when comparing results from different laboratories. With extremely
high air permeability spacer fabrics, tested to the ISO 11092 standard [170], the thermal
resistance is no longer dependent on the thickness due to air penetration, as shown in
Figure 5.

3.4. Materials vs. design


Crockford [75] recognised that the design of garments, particularly those designs which
encourage ventilation, is responsible for allowing relative comfort during hard work in

TTPR_A_845540.3d (TTPR)

14-02-2014 17:22

Textile Progress

165

Figure 5. The relationship between thermal resistance and thickness for fleece and spacer fabrics.

adverse conditions with impermeable fabrics. Other authors have recognised the significance of ventilation [171] and stressed the importance of clothing designs which enable
air penetration and subsequent heat removal. Lomax [106] states that WBFs should only
be regarded as one component in the overall moisture management system, and suggests
that factors such as moisture buffer layers, and ventilation features, are of equal, if not
more, importance.

3.5. Open-structure insulation


Particularly where large thermal gradients exist between the body and the environment,
heat transfer by radiation can be significant in foam structures [141] and also low-density
felted and knitted fabrics [172]. In thermal protective clothing, the insulating properties
of a fabric diminish with increasing air permeability [172,173]. Some innovations have
attempted to take advantage of this property (air flow dependence) of low-density insulation. For example, Spink [174] has patented a garment based on a specially shaped lowdensity polyurethane (PU) foam mid layer and a ventilated air-impermeable shell. In a
product evaluation for Salomon, a 3D mesh insulation top was tested, but participants felt
cold because the outer garment was permeable to air [175]. Krel et al. [176] found the
flocked fabrics allow greater heat fluxes and felt that this may, if incorporated into sports
clothing, increase labour productivity and influence the psychological and physiological
thermoregulation system, but did not test these claims. Recently, Helly Hansen has
released a fleece garment designed to increase ventilation, whereby the fleece fabric has
circular holes of different sizes removed from the fabric [177].

3.6. Synthetic and down insulation


As with base-layer garments there is considerable debate about the various pros and cons
of different types of insulation. As mentioned in the introduction, the dry thermal resistance of different materials differs very little and so of more importance is the way in
which different fibres interact with water, and practical considerations such as weight and
thickness required with different fibres. The different effects of water are the heat of sorption (which in Section 2 has been shown to be of little physiological significance),

TTPR_A_845540.3d (TTPR)

166

14-02-2014 17:22

M.P. Morrissey and R.M. Rossi

absorbed water decreasing insulation (thought to be higher in the case of natural fibres
coupled with slow drying times) and the effect of water on the structure of the insulation
(i.e. collapsing in the case of down insulation). It is commonly stated that synthetic insulation does not absorb water but in reality the situation is more complicated and deserves
closer inspection. The interaction depends on how the fabric is wetted: if synthetic and
down insulation are submerged in water and compressed, then naturally the inter-fibre
voids become filled with water. If the insulation is threatened with wetting by rain, then
the dependence is on the waterproofness or water repellence of the outer layer. Evaporated sweat may also condense inside the insulation. In each case, the primary difference
is that synthetic insulation retains its loft when wet, whereas the structure of down collapses and the down fibres can clump together [178]. In this case, the synthetic insulation
will retain more of its insulation than the down. There are also differences in the durability characteristics of down and synthetic insulation. Whilst the durability of down is
excellent in terms of repeated recovery from compression (resilience) [179,180], the durability as a garment is perhaps vulnerable, since in the case of the outer fabric of an insulated garment being damaged, the fibres are free to escape, whereas the synthetic batting
is not. Another consideration is the weight and thickness requirements of jackets made
from synthetic or down insulation: down provides better insulation per mass and synthetic
better insulation per thickness (due to the radiation-blocking density) [178]. Therefore, a
garment required to provide 3 Clo of insulation could be 1 mm thinner if made from synthetic insulation, a relatively small difference probably unnoticeable to the wearer,
whereas the down garment filling would require approximately half the mass (200 vs.
400 g, calculated from data in [181]).
A further detail regarding down insulation concerns down fill power, a measure of
the loft (volume occupied by a certain mass of down, usually given in cubic inches per
ounce) of a down product [182]. There are a number of different standards for measuring
loft, the two most pertinent being the US Federal test and the standard Lorch test, which
is approved by the International Down and Feather Testing Laboratory (IDFL): the IDFL
test yields results approximately 4% more conservative than the US test [183]. In a practical sense, higher fill power results in a lighter garment, as less down is required to create
the same loft, and therefore the same, or similar, insulation. Currently, PHD Mountain
Software offers 900 fill power down, and Patagonias Encapsil plasma treated water resistant down is claimed to have a fill power of 1000. Another topic is down overstuffing,
i.e. adding more down than necessary based on its fill power and the volume of garment
necessary to fill. Whilst this probably decreases the insulation per mass, overstuffing may
be useful to minimise the chances of dead spots where no down is present, and in maintaining loft in areas where garments are compressed.
3.7. Conclusions and future trends
This section has addressed insulation, which functions by trapping still air layers. Due to
this fact, when body movement or wind changes the characteristics of these air layers, the
clothing insulation also changes. By considering the separate mechanisms by which these
reductions in clothing insulation occur, clothing designers may be able to create clothing
with flexible insulation which can be adjusted to suit metabolic heat output, and avoid
unwanted heat loss due to wind. The wearer can also effect large changes in clothing insulation by using clothing openings.
The insulation per mass provided by down is difficult to improve upon, as its fibre
radius and fractal structure, studied with numerical modelling and other approaches,

TTPR_A_845540.3d (TTPR)

14-02-2014 17:22

Textile Progress

167

appear to be optimal [184187]. However, many innovations have recently emerged to


improve its water repellence. Some examples are Dri down, Down Tek and Encapsil
the latter being a plasma-applied coating. There is also a focus on improving the air permeability of insulation, e.g. Polartec Alpha insulation and the aforementioned Helly
Hansen H20 flow, a fleece with die or laser cut holes. Gibson et al. [188] found that
although nanofibres did not seem useful for high loft thermal insulation, they may play a
role in improving the performance of hybrid insulation with high bulk density. Plasmadeposited metal layers may be useful in reducing radiative heat loss in low-density insulation with better durability and negligible impact on evaporative resistance as was the case
with earlier reflective coatings [189191]. Aerogels have very low thermal conductivity,
but it is difficult to incorporate them into fabrics and clothing, as in order to overcome
their brittle qualities they must be supported by fibres, and the amount of fibres required
means that they paradoxically carry a weight penalty compared to other insulating materials, and cannot achieve comparable thermal conductivity at low bulk densities [188].

4. The shell layer


The role of the shell garment is to protect the inner layers from the ingress of wind and
rain, and provide mechanical protection to the whole clothing system. It should simultaneously allow the transfer of water vapour (evaporated sweat) from the clothing microclimate to the environment. Such fabrics are known as WBFs. WBFs are typically divided
into four categories: tightly woven fabrics, microporous, hydrophilic and bi-component
WBFs.

4.1. Waterproof breathable fabrics


4.1.1. Tightly woven WBFs
Tightly woven fabrics can be made from natural or synthetic fibres. The fabric structure is
responsible for waterproofness. For example, in the case of Ventile, large pores are avoided
by minimising crimp in the weft and using double-warp yarns [192]. The cotton fibres of
Ventile swell when wetted or exposed to high humidity, decreasing the inter-yarn pores
from 10 mm to 3 mm [193]; in this sense, Ventile can be considered a smart fabric, since
waterproofness is improved in conditions when it is required, and air permeability and
breathability improve when waterproofness is not required [194,195]. Tightly woven fabrics of synthetic fibres do not display this property but still impart water resistance.

4.1.2. Microporous WBFs traditional


Laminated or coated microporous WBFs are generally created from polymers such as polytetrafluoroethylene (PTFE) or PU; the microporous structure can be created via mechanical
fibrillation (bi-axial stretching at high temperature), wet cast coagulation (exposing viscous
PU to steam) or water vapour and salt dissolution (salt crystals added and subsequently
removed to create pores) [196]. Such films or coatings feature one to two billion pores per
square centimetre ranging in size from 0.1 mm to 3 mm [197]. Because water vapour molecules have diameters of 0.00004 mm, whereas liquid water droplets have diameters of at
least 100 mm, the complex passageways formed by these pores act as a liquid water filter
whilst still allowing the passage of water vapour. Because microporous coatings or films
can be compromised by water contaminated by detergents, sweat resides or salt water

TTPR_A_845540.3d (TTPR)

168

14-02-2014 17:22

M.P. Morrissey and R.M. Rossi

(lowering the surface tension and allowing the water to make a lower contact angle with the
porous network), many are coated with a solid hydrophilic layer [106,198] (bi-component
WBFs). A novel method of preventing this effect led to the event membrane [199].
Protecting the PTFE membrane without a hydrophilic coating was achieved by coating the
structure with a hydrophobic and oleophobic fluoropolymer; to achieve this, a wetting
agent is required to ensure the coating has a suitably low surface tension to wet and enter
the pores of the membrane, then a heating process evaporates the wetting agent.
4.1.3. Microporous WBFs nanofibre
Recently a new category of microporous membranes or coatings have emerged: these are
electrospun nanofibres. Garments based on electronspun nanowebs are already commercially available in the form of Polartec NeoShell : this is manufactured by Finetex, Inc.,
using a patented procedure [200]. In one study examining a number of bespoke and commercially available nanofibre-based WBFs, it was found that the Finetex, Inc., mass produced nanofibre web had the best combination of waterproofness and breathability [201].
Researchers have already conducted laboratory tests and human subject trials with microporous nanoweb WBFs: Lee and Obendorf [202] laminated electrospun nanofibres and
conventional spunbond non-wovens and found that the penetration of liquid water was
reduced whilst retaining a good air permeability in comparison to currently available protective fabrics, and with no significant reduction in moisture vapour transport. This air
permeability may improve comfort [203]. Lee and Obendorf observed that electrospun
nanofibre webs have pores that are larger than those in conventional microporous membranes but smaller than those in conventional spunbond non-wovens used for protective
clothing [204]. Gibson et al. [205] propose that distortion effects on elastomeric-based
nanofibre webs may significantly affect the transport behaviour. Bagherzadeh et al. [201]
laminated electrospun nanofibre webs and woven fabrics, and found that electrospun
nanofibre mats also have improved water vapour transfer properties compared to conventional PTFE-based WBFs and acceptable (i.e. lower) waterproofness. Ahn et al. [206]
also conducted human wearer trials and found that whilst the water resistance of nanofibre
webs was indeed lower, it is probably sufficient for protection from rain. In their wearer
trials, they found evidence that the nanofibre WBFs provide more comfortable conditions
in dry weather (evidenced by lower clothing microclimate absolute humidity), but in wet
weather (simulated rain) there was no difference between conventional PTFE-based
WBFs and nanofibre WBFs. Rietveld [207,208] conducted field trials comparing conventional WBFs and newly commercially available nanofibre-based WBFs. The clothing
microclimate humidity was recorded, and it was concluded that all types of WBFs accumulated condensation and that the new air-permeable membranes, event and NeoShell , represent a small but measureable advantage in terms of microclimate humidity
and drying time. Rietveld perceived ventilation to be of more importance than fabric
breathability, and that wearing a rucksack compromised the performance of all the garments on test.
4.1.4. Solid hydrophilic coating or laminated membrane
Hydrophilic WBFs are based on block co-polymers usually comprising PU or PES and
polyethylene oxide (PEO) or polyethylene glycol (PEG) (often 40%60% weight for
weight, respectively) [106,209]. They are essentially non-porous (in the conventional
sense) and impermeable to air. Hydrophilic coatings are not monolithic but instead feature

TTPR_A_845540.3d (TTPR)

14-02-2014 17:22

Textile Progress

169

distinct functional layers defined as the polyethylene glycol tie coat which lies next to
the face fabric, the second coat which lies in the middle and finally the top coat. The
tie coat must be highly flexible and very breathable in order for desirable fabric handle
and moisture vapour transmission (MVT) through the fabric interstices. It must also
adhere sufficiently to the yarns of the fabric; these characteristics are usually obtained by
manipulating the degree of polymer cross linking and increasing the PEO quantity. The
top coat is engineered to be durable, thermoplastic for seam taping and wettable in order
to maximise absorption of water vapour. This is achieved by using lower quantities of
PEO and avoiding cross linking.
Hydrophilic films and coatings swell in conditions of high water vapour pressure,
becoming more breathable as a result. They can thus be described as smart or intelligent
textile products. At low water vapour pressure, hard PU segments agglomerate together
via inter- and intra-molecular hydrogen bonding, effectively forming a constraining mesh
around the soft PEO segment. As water vapour pressure increases, the PEOPU membrane swells, creating a more PEO-rich surface cable of increased MVT.
The process of MVT through these coatings has been described as one of stepping
stones or molecular wicking [210,211]. The film functions because water molecules
have a stronger affinity for the functional PEO molecules than to other water molecules
(3438
kJ mol1 vs. 1923 kJ mol1). The similarity between the adjacent OO distance

(2.8 A) in the PEO and water molecule clusters accounts for the high permeability of
PUPEO films and coatings [106]. Weak hydrogen bonds are formed between the electronegative oxygen of the water molecules and the electropositive outer of the PEO (in some
instances PEG): this is described as a hydration sheath [106]. Compared to covalent
bonds between oxygen and hydrogen (OH), hydrogen bonds (HH) are only 4% as
strong (492 kJ mol1 vs. 23 kJ mol1). Thus, these bonds are easily broken and water
molecules are displaced along the chain due to the water vapour pressure gradient created
by the temperature and humidity on either side of the hydrophilic layer.
4.1.5. Bi-component WBFs
Bi-component WBF fabrics are a combination of the previously described microporous
and hydrophilic fabrics, usually laminated to a tightly woven face fabric. GORE-TEX is
the most well known example of a bi-component WBF. In this case, the hydrophilic coating partially fills the microporous structure [198].

4.2. Performance of WBFs and their differences


4.2.1. Ficks law and WBFs
Fick first proposed the laws of diffusion based on Fouriers laws of heat transfer [210].
Rearranged to give the dimensions of most MVT tests, Ficks law states that the flux
of a substance Q per unit area A per time T is proportional to the concentration gradient
(c1 c2):
Q
c1  c2
D
;
At
l
where D is the diffusion coefficient and l is the thickness of the fabric, membrane or coating. This equation is only appropriate for steady-state diffusion where D is constant, such
as with a microporous PTFE membrane or tightly woven synthetic fabric. Hygroscopic

TTPR_A_845540.3d (TTPR)

170

14-02-2014 17:22

M.P. Morrissey and R.M. Rossi

natural fibres such as cotton used in Ventile, and hydrophilic coatings, interact with the
diffusing water vapour and the diffusion coefficient D can vary.
Due to this dependency on the water vapour concentration gradient, c1c2, the water
vapour transfer rate of 10,000 g m2 d1, which is required at high exercise intensities
(up to 1000 W), is unobtainable at realistic pressure gradients (100300 Pa) [106]. There
is a large discrepancy between such real-life pressure gradients and those created in methods to test the breathability of fabrics: 3093, 3168 and 818 Pa for ISO 11092, ISO 15946
and ISO 8096, respectively.
4.2.2. Effects of temperature on diffusion resistance
Different types of membrane are affected by low temperature and rain in different ways.
Hydrophilic membranes show greater water uptake at lower temperatures: the magnitude
of this effect is dependent on the PEO content and chain length [106]. There is some conflicting evidence regarding the performance of hydrophilic WBFs in cold conditions in
the literature. Osczevski found that the diffusion resistance of the hydrophilic component
of GORE-TEX increased greatly at sub-zero temperatures such that MVT was reduced
to just a few per cent of its room temperature value [212]. Umbach and Bartels conducted
wearer trials showing that moisture accumulation was significantly lower in hydrophilic
garments than in impermeable garments, which somewhat contradicted Osczevskis findings [213]. Gibson elucidated this apparent contraction by independently controlling temperature and humidity gradients, using a dynamic moisture permeation cell and showing
that the diffusion resistance of hydrophilic coatings is far more affected by humidity than
temperature, and that Osczevskis misinterpretation was due to the inability of his apparatus to control humidity on either side of the sample [214]. Despite this, decreases in MVT
at low temperatures are likely due to the relationship between saturated vapour pressure
and temperature, rather than due to changes in polymer behaviour. Microporous WBFs
show little temperature dependence in terms of changes in polymer behaviour [214].
4.2.3. Effects of low temperature and rain
The main concern regarding the use of WBFs in cold or rainy conditions is the formation
of condensation on the inner surface of the WBF. Whereas microporous WBFs can
become physically blocked by rain, rendering them impermeable [192], solid hydrophilic
and bi-component WBFs continue breathing, provided there is a favourable water
vapour pressure gradient. Both Gretton [215] and Rossi et al. [216] found that condensation accumulation in rain, and at low temperatures, is largest in microporous WBFs, followed by hydrophilic and bi-component WBFs. Gretton also found the same ranking
in field trials [194]. In simulated conditions of rain at 5 C, MVT rates of 0, 1050 and
1800 g m2 d1 were measured for microporous, hydrophilic and bi-component WBFs,
respectively [215]. The condensation accumulation was proportional to that measured in
wearer trials, and inversely proportional to the temperature gradient across the shell fabric. The poor performance of microporous membranes is attributed either to physical
blocking of pores by rain or by the condensation of moisture within the microporous
structure itself [216]. Since the hydrophilic component of pure hydrophilic and bi-component membranes continues to breathe, Gretton attributed the improved performance of
bi-component WBFs to the trapped air layer in the PTFE structure which maintained a
measurably larger temperature difference across the fabric, inhibiting condensation [215].
More simply, lower condensation would also be expected in fabrics with low evaporative

TTPR_A_845540.3d (TTPR)

14-02-2014 17:22

Textile Progress

171

resistance [217,218]. It has also been suggested that a tricot layer with good transverse
wicking properties attached to the shell inner will also promote re-evaporation through
the WBF as the condensation is spread over, and can evaporate from, a larger area.
The durable water repellent (DWR) coating of the shell layer is also extremely important in minimising condensation accumulation [15,215,219]. With a poor DWR, the fabric
wets out and the difference in surface temperature between the inside and outside of the
shell garment quickly decreases [215,219]. The increase in conductivity related to this
increases condensation accumulation, thus causing a decrease in MVT. These effects continue in a vicious circle of events, leading to increasing condensation accumulation and
poor MVT. Whilst wind without precipitation tends to increase MVT by disturbing the
boundary layer and maintaining a favourable moisture vapour pressure gradient, wind in
combination with rain increases condensation as the shell fabric is cooled, leading to
increased condensation accumulation [220,221].
4.3. Conclusions and future trends
The various types of WBF offer different benefits and have their own unique drawbacks.
The poor performance of microporous membranes in terms of condensation accumulation
and limited breathability in rain has likely been addressed by the new PTFE-based event
membrane and electrospun nanofibre web NeoShell . In many conditions, the air permeability of these WBFs may present an advantage in terms of thermal comfort, but in
very windy conditions, continuous, air-impermeable coatings such as those used in
SympaTex and GORE-TEX may prove advantageous in terms of conserving body
heat. The durability of bi-component WBFs may also be greater due to their composite
construction. One area where significant improvements may be made is in super hydrophobic coatings. Silicone nanofilament coatings may present significant advantages in
this area [222,223], and future biomimetic technology may be able to self-heal using similar self-assembly mechanisms to those demonstrated by the original super hydrophobic
material, the lotus leaf [224,225].

5. Further discussion
5.1. The complete clothing system: interaction
So far it has been attempted to discuss the separate layers of a typical three-layer clothing
system independently. Of course, when these layers are worn together, they interact. The
simplest example of this is that when fabrics are worn as an ensemble, the total clothing
insulation is more than the sum of the thermal resistance of each material layer: Havenith
et al. [112] estimate that for a two-layer clothing ensemble, more than 60% of the total
insulation can be attributed to air layers between the skinfabric and fabricfabric layers,
and less than 40% can be attributed to the clothing layers themselves. This effect is also
valid for evaporative resistance. Condensation accumulation in clothing can also be modified by changing the order or fabric layers, without changing the actual fabric layers
themselves: Yoo and Kim [226] found that by placing an insulating fabric next to the
outer shell layer, rather than an air gap, condensation accumulation was reduced. This
effect may explain data suggesting that more condensation accumulates in membrane vs.
non-membrane soft-shell garments, the latter of which often feature an insulating pile or
fleece layer attached to the tightly woven shell [15]. Liquid water in the clothing system,
be it condensed sweat, non-evaporated sweat or rain, can be transferred between the

TTPR_A_845540.3d (TTPR)

172

14-02-2014 17:22

M.P. Morrissey and R.M. Rossi

clothing layers [227]. The amount of transfer has been shown to be partly dependent on
the pressure applied to the two fabrics in contact, which could, for example, be created by
wearing a rucksack [228]. When the pressure applied to the two fabrics is low, there is little or no transfer of liquid water, and at higher pressures there are no voids in the fabric
layer for water to enter. Additionally, the orientation of single-sided fleece affects liquid
transfer between layers: when the pile side of the dry fabric contacted either side of the
wet fabric, there was no liquid transfer, and when the backing side of the dry fabric contacted either side of the wet fabric, there was. The larger amount of liquid water transfer
observed with backing-to-backing orientation is due to the small capillary size, which
effect was also observed by Osczevski and Crow [37] who found that water uptake was
related to the pore size. In conclusion, in addition to considering the structure and properties of the individual layers, the designer of a clothing system must consider how these
layers will interact. Especially important is to try and avoid the problem of subsequent
layers attenuating the properties of the underwear and insulation: Laing et al. [79] showed
that if two base layers with measurably different properties are combined with another
mid or shell layer, it is possible for there to be no measurable difference in the new assembly. Numerical thermal manikin and thermo-physiological models (e.g. [229231]), and
computational fluid dynamics software (e.g. [232]), may prove to be increasingly useful
for designers and textile engineers attempting to predict the behaviour of complete clothing systems.
5.2. Test methods for the appraisal of clothing
Both Umbach [233] and Goldman [102] have developed similar multi-layer systems for
the assessment of clothing system performance. At the first level, a large number of test
methods are conducted with fabric samples to ascertain the basic structural parameters
and properties of fabrics. These methods are quick, precise and logistically simple. The
second level is the first test of garments, using a thermal manikin to provide objective and
precise data. The final level tests garments using human subjects either in controlled laboratory conditions or in real-use. At this level, testing becomes more logistically complex,
and explaining the behaviour of the clothing system based on fundamental principles
becomes more difficult, due to the large number of physiological, environmental and textile variables. However, this testing does provide important information about the functionality of the clothing in target-use conditions. Goldman states that with each
incremental level of testing, yield of scientific information and reproducibility decrease,
and the cost and number of potentially confounding variables increase.
Most test methods used to determine characteristics such as thickness, mass per unit
area and thermal properties are used without much discussion or debate as to the validity
of the method. As discussed in Section 3.5, it is only extremely open fabrics that are
highly influenced by air flow, and allow unusually large heat transfer by radiation, that
cause discrepancies in thermal resistance determination. However, the determination of
the moisture vapour permeability of fabrics has caused substantial controversy, especially when the use of different methods has been selected to suit inflated marketing
claims [106].
The main reason that these tests have caused so much controversy lies in the simple
fact that the water vapour pressure gradients created across the WBF are far larger than
usually present in real-use. As previously mentioned in Section 4, the water vapour pressure gradient in real-life is only a few hundred Pa, as opposed to a few thousand in most
test methods [106]. The inverted-cup method, BS EN ISO 15496, produces a water vapour

TTPR_A_845540.3d (TTPR)

14-02-2014 17:22

Textile Progress

173

pressure gradient of 2168 Pa using desiccants, and although being widely discredited in
terms of its relation to real-life, is fast and therefore a useful quality control measure. Only
when the results are applied to real-life scenarios do problems arise. The sweating-guarded
hot plate method [170] works on the principle of a textile being placed on a heated plate
saturated water; the test is isothermal (35 C), but the environment is maintained at 40%
RH and the saturated plate assumed to be 100% RH. Because of the high temperature in
the hot plate method, and low humidity next to the desiccant in the inverted-cup method,
both methods underrate hydrophilic membranes which are engineered to function optimally at low temperatures, and show greater permeability in high humidity or when wet
(as with condensation in real-life). This has led to the hot plate method being described as
utterly meaningless for foul-weather garments worn in real-life [106].

5.3. Smart materials


Smart materials can be defined as materials that exhibit relatively large and dramatic
physical, chemical and/or biological changes in response to external stimuli [234].
Smart textiles has been described as one of the great catch phases of 21st century
textiles, and due to the limited number of commercial products, it has been suggested
that the research in this area is in danger of losing. . . credibility [235]. Though progress has perhaps been slower than the optimistic claims of some researchers and popular science publications might suggest, there is still some promise of smart materials
improving the performance and functionality of protective clothing. For example,
advances in energy harvesting from body movement and warmth continue to be made
using piezoelectric and thermoelectric textiles, respectively [236,237]. Such technology may allow the partial powering of mobile phone or GPS devices which are popular with outdoor enthusiasts. Piezoelectric textiles can also be used to monitor vital
body signals such as respiration and pulse rate, and can act as input keys for textile
keyboards or buzzer alerts [236]. Shape memory polymers (SMPs) have also received
attention from researchers interested in textile applications. Potential applications
with SMPs range from temperature-dependent, water vapour permeable membranes to
garments that can roll up their own sleeves [238,239]. SMPs can also be used to construct insulation that changes its thickness depending on the ambient temperature;
while this change in insulation is not fully sufficient to account for all changes in the
ambient temperature, it has been shown to extend the useful temperature range of an
item of clothing [240]. For outdoor activities requiring body armour, such as mountain
biking or extreme snow sports, the use of magnetostrictive materials in combination
with fabrics, allowing smart or adjustable armour stiffness and impact dissipation, is
a possibility [241]. Chromogenic and halochromic textiles have seen applications in
industrial settings, where the colour-changing fabric may alert the wearer to hazards
such as carbon monoxide or dangerous acid vapours [242]. While applications in outdoor clothing are less clear, such chromogenic textiles might offer improved aesthetic
appeal to consumers [243]. Finally, a silicone-based, self-healing textile is now a commercial reality, potentially improving the longevity of outdoor equipment [244].

6. Summary and conclusions


 The material properties and textile structures comprising an outdoor clothing system can have an effect on human safety, comfort and performance. This literature

TTPR_A_845540.3d (TTPR)

14-02-2014 17:22

174

M.P. Morrissey and R.M. Rossi

review has shown that the effect of fibre type and textile structure on the wearer can
be highly dependent on the environmental conditions and type of activity
conducted.
 The layers of a clothing system interact and behave differently as a system from
that might be expected from their individual properties. Therefore, designers of
clothing systems face a considerable challenge to ensure not only that the most
appropriate materials and structure are selected for each layer, but that that as a system, these layers interact desirably: i.e. they do not attenuate or nullify desired
properties of the other layers.
 Test methods must be carefully considered to avoid producing misleading
information.
 Many other factors which have not been addressed in this review must be considered, such as clothing fit and ergonomic factors, clothing weight [245], design and
styling. Only by considering all these factors can clothing for the outdoors be optimised, ensuring widespread, safe and comfortable participation in outdoor activities.

Acknowledgements
The first author would like to acknowledge Mr Dave Brook at the University of Leeds for his brilliant lectures which were instrumental in inspiring the author to undertake research in technical textiles. Credit is also due to Dr Mark Taylor, also at the University of Leeds, for years of advice,
discussion and debate about this topic. Thanks also to Dr Robert Lomax, Dr Rene Rossi and the
Editor-In-Chief Professor Richard Murray for encouraging and enabling the author to write and
publish this literature review.

References
[1] J.M. Bowker, D.B. English and H.K. Cordell, Projections of outdoor recreation participation
to 2050, in Outdoor Recreation in American Life: A National Assessment of Demand and
Supply Trends, H. Ken Cordell, ed., Sagamore Publishing, Champaign, IL, 1999, p.323.
[2] S. De Vries, R.A. Verheij, P. Groenewegen and P. Spreeuwenberg, Environ. Plan. A 35(10)
(2003) p.1717.
[3] G. Godbey, Outdoor Recreation, Health and Wellness: Understanding and Enhancing the
Relationship, RFF Press, Washington, DC, 2009. Available at http://ssrn.com/
abstract=1408694.
[4] C. Maller, M. Townsend, A. Pryor, P. Brown and L. St Leger, Health Promot. Int. 21(1)
(2006), p.45.
[5] H. Frumkin, Am. J. Prev. Med. 20(3) (2001) p.234.
[6] P.N. Ainslie, I.T. Campbell, K.N. Frayn, S.M. Humphreys, D.P.M. Maclaren and T. Reilly, J.
Appl. Physiol. 92(1) (2002) p.179.
[7] P.N. Ainslie, I.T. Campbell, J.P. Lambert, D.P.M. MacLaren and T. Reilly, Sports Med. 35(7)
(2005) p.619.
[8] R. Laing and G. Sleivert, Text. Prog. 32(2) (2009) p.1.
[9] W. McArdle, F. Katch and V. Katch, Energy for physical activity, in Exercise Physiology:
Nutrition, Energy and Human Performance, Lippincott Williams & Wilkins, Baltimore, MA,
2009, p.107.
[10] R. Nave, HyperPhysics. Available at http://hyperphysics.phy-astr.gsu.edu/hbase/hframe.
html.
[11] W. McArdle, F. Katch and V. Katch, Exercise performance and environmental stress, in Exercise Physiology: Nutrition, Energy and Human Performance, Lippincott Williams &
Wilkins, Baltimore, MA, 2009, p.616.
[12] U. Hipler and P. Elsner, Biofunctional Textiles and the Skin, S. Karger AG, Basel, 2006.
[13] M. Parsons and M. Rose, Invisible on Everest: Innovation and the Gear Makers, Old City
Publishing, Philadelphia, PA, 2002.

TTPR_A_845540.3d (TTPR)

14-02-2014 17:22

Textile Progress

175

[14] J.E. Ruckman, J. Fashion Mark. Manage. 9(1) (2005) p.122.


[15] K. Stevens, Thermophysiological Comfort and Water Resistant Protection in Soft Shell Protective Garments, University of Leeds, Leeds, 2008.
[16] Q. Chen, J. Fan, M.K. Sarkar and K. Bal, Text. Res. J. 81(10) (2010) p.1039.
[17] E. Kissa, Text. Res. J. 66(10) (1996) p.660.
[18] A. Patnaik, R.S. Rengasamy, V.K. Kothari and A. Ghosh, Text. Prog. 38(1) (2006) p.1.
[19] C. Qing, Development of Plant Structured Knitted Fabrics, The Hong Kong Polytechnic
University, Hong Kong, 2010.
[20] Y. Zhang, H.P. Wang and Y.H. Chen, J. Appl. Polym. Sci. 102(2) (2006) p.1405.
[21] R.K. Varshney, V.K. Kothari and S. Dhamija, J. Text. Inst. 101(6) (2010) p.495.
[22] G. Fauland, P. Hofer, W. Nachbauer and T. Bechtold, Text. Res. J., 82(2) (2011) p.99.
[23] G. Supuren, N. Oglakcioglu, N. Ozdil and A. Marmarali, Text. Res. J. 81(13) (2011) p.1320.
[24] M.M. Hossain, A.S. Herrmann and D. Hegemann, Plasma Process. Polym. 3(3) (2006) p.299.
[25] A.B. Nyoni and D. Brook, Text. Res. J. 80(8) (2010) p.720.

[26] E. Oner,
H.G. Atasagun, A. Okur, A.R. Beden and G. Durur, J. Text. Inst. 104 (2013) p.699.
[27] W. Wardiningsih and O. Troynikov, J. Text. Inst. 103(1) (2012) p.89.
[28] E. Rodwell, E. Renbourn, J. Greenland and W. Kenchington, J. Text. Inst. Trans. 56(11)
(1965) p.T624.
[29] A.C. Burton and O.G. Edholm, Man in a Cold Environment. Physiological and Pathological
Effects of Exposure to Low Temperatures, Arnold, London, 1955.
[30] D.M. Kerslake, The Stress of Hot Environments, Cambridge University Press, Cambridge, 1972.
[31] G. Havenith, P. Brode, V. Candas, E. den Hartog, I. Holmer, K. Kuklane, H. Meinander, W.
Nocker, M. Richards and X. Wang, Evaporative cooling in protective clothing efficiency
in relation to distance from skin, in Proceedings of the 13th International Conference on
Environmental Ergonomics, University of Wollongong, Wollongong, New South Wales,
Australia, 2009, p.20.
[32] G. Havenith, P. Brode, E. den Hartog, K. Kuklane, I. Holmer, R.M. Rossi, M. Richards, B.
Farnworth and X. Wang, J. Appl. Physiol. 114(6) (2013) 778.
[33] F. Wang, S. Annaheim, M. Matthew and R.M. Rossi, Evaporative cooling efficiency of onelayer tight fitting sportswear: a sweating torso manikin study, in XV International Conference on Environmental Ergonomics, International Society for Environmental Ergonomics,
Loughborough, 2013, p.285.
[34] R. Jordan and M. Martin, Just Say No To Wicking: Non-Traditional Base Layers Based on a
Next-to-Skin Fishnet Model, 2012. Available at http://www.backpackinglight.com/cgi-bin/
backpackinglight/fishnet_base_layers.html.
[35] XBionic, X-bionic Catalogue, 2010. Available at http://www.x-bionic.com/.
[36] G. Havenith, Exogenous Dermatol. 1(5) (2002) p.221.
[37] R.M. Crow and R.J. Osczevski, Text. Res. J. 68(4) (1998) p.280.
[38] L. Fourt, A.M. Sookne, D. Frishman and M. Harris, Text. Res. J. 21(1) (1951) p.26.
[39] S. Yoo and R.L. Barker, Text. Res. J. 74(11) (2004) p.995.
[40] R.M. Laing, B.E. Niven, R.L. Barker and J. Porter, Text. Res. J. 77(3) (2007) p.165.
[41] C. Prahsarn, Text. Res. J. 75(4) (2005) p.346.
[42] E. Baussan, M. Bueno, R. Rossi and S. Derler, Text. Res. J. 83 (2012) p.836.
[43] C.P. Bogerd, I. Rechsteiner, B. Wust, R.M. Rossi and P.A. Br
uhwiler, Ann. Occup. Hyg. 55
(5) (2011) p.510.
[44] Z. Vokac, V. Kopje and P. Keul, Text. Res. J. 46 (1976) p.30.
[45] I. Holmer, Text. Res. J. 55(9) (1985) p.511.
[46] Y. Li, J.H. Keighley and I.F. Hampton, Ergonomics 31(11) (1988) p.1709.
[47] T.P. Gavin, J.P. Babington, C.A. Harms, M.E. Ardelt, D.A. Tanner and J.M. Stager, Med. Sci.
Sports Exerc. 33 (2001) p.2124.
[48] B.C. Roberts, T.M. Waller and M.P. Caine, Int. J. Sports Sci. Eng. 1(1) (2007) p.29.
[49] M.K. Bakkevig and R. Nielsen, Ergonomics 38(5) (1995) p.926.
[50] R. Anderson, Full Exposure: U.S. Army Data Helps Clear the View of Waterproof/breathable, 2004. Available at: http://www.insideoutdoor.com/documents/W-BJune04feature.pdf.
[51] A. Plante, B. Holcombe and L. Stephens, Text. Res. J. 65(5) (1995) p.293.
[52] Y. Li, A.M. Plante and B.V. Holcombe, Text. Res. J. 65(6) (1995) p.316.
[53] R. Niedermann and R. Rossi, Text. Res. J. 82(4) (2012) p.374.
[54] Y. Li, B.V Holcombe and F. Apcar, Text. Res. J. 62 (1992) p.619.

TTPR_A_845540.3d (TTPR)

176

14-02-2014 17:22

M.P. Morrissey and R.M. Rossi

[55] R. Fangueiro, P. Gonc alves, F. Soutinho and C. Freitas, Ind. J. Fibre Text. Res. 34 (2009)
p.315.
[56] K. Varga, U. Schadel, H. Nilsson, O. Persson and K.C. Schuster, FibreText. Eastern Europe
15 (2007) p.59.
[57] J.T. Nelbach and L. Herrington, Science 95(2467) (1942) p.387.
[58] P.A. Siple, Proc. Am. Philos. Soc. 89(1) (1945) p.200.
[59] B. Farnworth, Calculations of Heat and Vapour Transport in Clothing: Transient Effects in
Hygroscopic Materials (Ottawa Technical Note 80-2lIVERLIBNE5), Defence Research
Establishment, Ottawa, 1986.
[60] B. Farnworth, Text. Res. J. 56(11) (1986) p.653.
[61] L. Fengzhi and L. Yi, Model. Simul. Mater. Sci. Eng. 13(6) (2005) p.809.
[62] P. Gibson and M. Charmchi, J. Appl. Polym. Sci. 64(3) (1997) p.493505.
[63] M. Ha, H. Tokura and Y. Yamashita, Ergonomics 38(7) (1995) p.1445.
[64] J. Wickwire, P.A. Bishop, J.M. Green, M.T. Richardson, G.R. Lomax, C. Casaru, M. CurtherSmith and B. Doss, Int. J. Ind. Ergonom. 37(7) (2007) p.643.
[65] A. van den Heuvel, P. Kerry, J. van der Velde, M. Patterson and N. Taylor, Can undergarments be of benefit when working in protective clothing in hot environments?, in Proceedings
of the 13th International Conference on Environmental Ergonomics, John W. Castellani and
Thomas L Endrusick, eds., University of Wollongong, Wollongong, New South Wales, 2009,
p.35.
[66] I. Holmer, Text. Res. J. 55(9) (1985) p.511.
[67] R. Nielsen and T. Endrusick, The Role of Textile Material in Clothing on Thermoregulatory
Responses to Intermittent Exercise, US Army Research Institute of Environmental Medicine,
Natick, MA, 1988.
[68] D.C. Gavhed and I. Holmer, Eur. J. Appl. Physiol. Occup. Physiol. 73(6) (1996) p.573.
[69] M. Ha, Y. Yamashita and H. Tokura, Eur. J. Appl. Physiol. Occup. Physiol. 71(23) (1995)
p.266.
[70] A. Kwon, M. Kato, H. Kawamura, Y. Yanai and H. Tokura, Eur. J. Appl. Physiol. 78 (1998)
p.487.
[71] K. Tanaka, K. Hirata and Y. Kamata, Eur. J. Appl. Physiol. 84(12) (2001) p.69.
[72] R.M. Laing, S.T. Sims, C.a. Wilson, B.E. Niven and N.M. Cruthers, Ergonomics 51(4)
(2008) p.492.
[73] S.X. Wang, Y. Li, H. Tokura, J.Y. Hu, Y.X. Han, Y.L. Kwok and R.W. Au, Text. Res. J.
77(12) (2007) p.968.
[74] Y. Guo, Y. Li, H. Tokura, T. Wong, J. Chung, A.S. Wong, M.D.I. Gohel and P.H.M. Leung,
Text. Res. J. 78(12) (2008) p.1057.
[75] G.W. Crockford, M. Crowder and S.P. Prestidge, Occup. Environ. Med. 29(4) (1972) p.378.
[76] J.O. Kim and S.M. Spivak, Text. Res. J. 64(2) (1994) p.112.
[77] J.C. Barnes and B.V. Holcombe, Text. Res. J. 66(12) (1996) p.777.
[78] I.M. Stuart, A.M. Schneider and T.R. Turner, Text. Res. J. 59(6) (1989) p.324.
[79] R.M. Laing, B.A. MacRae, C.A. Wilson and B.E. Niven, Text. Res. J. 81(17) (2011) p.1828.
[80] B. Farnworth and P. Dolhan, Text. Res. J. 55(10) (1985) p.627.
[81] W. Lotens and G. Havenith, Ergonomics 38(6) (1995) p.1092.
[82] M. Naebe and B.A. McGregor, J. Text. Inst. 83 (2013) p.1.
[83] R. Nielsen and T. Endrusick, Eur. J. Appl. Physiol. 60 (1990) p.15.
[84] B. Roberts and G. Bertram, Handbook on Clothing and Equipment Required in Cold
Climates, The War Office, London, 1941.
[85] G.F. Fonseca, Text. Res. J. 40(6) (1970) p.553.
[86] R.F. Goldman and B. Kampmann, Handbook on Clothing: Biomedical Effects of Military
Clothing and Equipment Systems, International Society for Environmental Ergonomics,
Loughborough, 2007.
[87] H. Ueda, Y. Inoue, M. Matsudaira, T. Araki, and G. Havenith, Int. J. Cloth. Sci. Technol.
18(4) (2006) p.225.
[88] M.K. Bakkevig and R. Nielsen, Ergonomics 37(8) (1994) p.1375.
[89] M. Critchley, BMJ. 174 (1945) p.173.
[90] A.H. Woodcock and T.E. Dee, Wet - cold I: Effect of Moisture on Transfer of Heat Through
Insulating Materials, Department of the Army Office of the Quartermaster General, Fort
Lee, VA, 1950.

TTPR_A_845540.3d (TTPR)

14-02-2014 17:22

Textile Progress

177

[91] L. Hes and M. de Araujo, Text. Res. J. 80(14) (2010) p.1488.


[92] D.M. Kerslake, Ergonomics 31(7) (1988) p.1009.
[93] L. Hes, P. Offermann and I. Dvorakova, The effect of underwear on thermal contact feeling
caused by dressing up and wearing of garments, in Tecnitex Autex Conference, 2001.
Available at https://www.google.com/url?sa=t&rct=j&q=&esrc=s&source=web&cd=1&ved=
0CCsQFjAA&url=http%3A%2F%2Fcentrum.tul.cz%2Fcentrum%2Fpublikace%2Fpristroje%
2F%255B21%255DHesOfferDvor.doc&ei=IqBKUsCkDuWJ4ATgo4BA&usg=AFQjCNFlrm
LRRexu6UO-25ME5sMtdVfdmg&sig2=QTESoYShdDDmmhFxMODriw&bvm=bv. 53371865,
d.bGE. Accessed 1st October 2013.
[94] J.H. Andreen, J.W. Gibson and O.C. Wetmore, Text. Res. J. 23(1) (1953) p.11.
[95] R. Nielsen, D.C. Gavhed and H. Nilsson, Ergonomics 32(12) (1989) p.1581.
[96] N.J. Shirtcliffe, G. McHale, M.I. Newton, C.C. Perry and P. Roach, Chem. Commun. (25)
(2005) p.3135.
[97] F. Xia, L. Feng, S. Wang, T. Sun, W. Song, W. Jiang and L. Jiang, Adv. Mater. 18(4) (2006)
p.432.
[98] T. Sun, G. Wang, L. Feng, B. Liu, Y. Ma, L. Jiang and D. Zhu, Angew. Chem. (Int. ed. in
Engl.) 43(3) (2004) p.357.
[99] K. Ledward, Natural vs synthetic base layers. Available at http://www.klets.co.uk/Natural vs
Synthetics Report.pdf.
[100] W. Lotens, Heat Transfer From Humans Wearing Clothing, Technical University Delft, Delft,
1993.
[101] I. Holmer, Textiles for protection against cold, in Textiles for Protection, R. Scott, ed., Woodhead Publishing, Cambridge, 2005, p.378.
[102] R. Goldman, Thermal Manikins, Their Origins and Role, in Sixth International Thermal
Manikin and Modelling Meeting, J. Fan, ed., The Hong Kong Polytechnic University, Hung
Hom, Kowloon, Hong Kong, 2006, p.3.
[103] S. Backer, Text. Res. J., 18(11) (1948) p.650.
[104] B. Jones and E. McCullough, A Comprehensive Data Base For Estimating Clothing Insulation, Kansas State University Institute For Environmental Research, Manhattan, KS, 1984.
[105] S.K. Obendorf and J.P. Smith, Text. Res. J. 56(11) (1986) p.691.
[106] G.R. Lomax, J. Mater. Chem. 17(27) (2007), p.2775.
[107] C. Winslow, A. Gagge and L. Herrington, Am. J. Physiol. 127 (1938) p.505.
[108] W. H. Rees, J. Text. Inst. 37 (1946) p.132.
[109] L. Fourt and N. Hollies, Clothing Comfort and Function, Marcel Dekker, New York, 1970.
[110] G.S. Campbell, A.J. Mcarthur and J.L. Monteith, Bound.-Layer Meteorol. 18 (1980) p.485.
[111] B. Givoni and R.F. Goldman, J. Appl. Physiol. 6 (1972) p.812.
[112] G. Havenith, R. Heus and W. Lotens, Ergonomics 33(1) (1990) p.67.
[113] H. Anttonen, E. Hiltunen, I. Holmer and G. Ohlsson, The effect of cold, wind and movements
on clothing insulation, in Problems with Cold Work, Ingvar Holmer and Kalev Kuklane, eds.,
Arbetslivsinstitutet, Solna, Sweden, 1997, p.77.
[114] H. Nilsson, I. Holmer, G. Ohlsson and H. Anttonen, Clothing insulation at high wind speed,
in Problems with Cold Work, Ingvar Holmer and Kalev Kuklane, eds., Arbetslivsinstitutet,
Solna, Sweden, 1997, p.114.
[115] I. Holmer, H. Nilsson, G. Havenith and K. Parsons, Ann. Occup. Hyg. 43(5) (1999) p.329.
[116] R. Rossi, M. Weder, R. Gross and F. Kausch, Influence of air permeability on thermal and
moisture transport through clothing, in Ergonomics of Protective Clothing (Proceedings of
NOKOBETEF 6 and 1st European Conference on Protective Clothing), Ingvar Holmer and
Kalev Kuklane, eds., Arbetslivsinstitutet, Solna, Sweden, 2000, p.12.
[117] H. Nilsson, H. Anttonen and I. Holmer, New algorithms for prediction of wind effects on cold
proteective clothing, in Ergonomics of Protective Clothing (Proceedings of NOKOBETEF 6
and 1st European Conference on Protective Clothing), Ingvar Holmer and Kalev Kuklane,
eds., Arbetslivsinstitutet, Solna, Sweden, 2000, p.17.
[118] I. Holmer, H. Nilsson and H. Anttonen, Prediction of wind effects on cold protective clothing,
in Blowing Hot and Cold: Protecting Against Climatic Extremes, NATO Science and Technology Organization, Neuilly-sur-Seine, 2001, p.8.
[119] G. Havenith and H.O. Nilsson, Eur. J. Appl. Physiol. 92(6) (2004) p.636.
[120] L.M. Bouskill, G. Havenith, K. Kuklane, K.C. Parsons and W.R. Withey, AIHA J. (Fairfax,
VA) 63(3) (2002) p.262.

TTPR_A_845540.3d (TTPR)

178

14-02-2014 17:22

M.P. Morrissey and R.M. Rossi

[121] J.J. Vogt, J.P. Meyer, V. Candas, J.P. Libert and J.C. Sagot, Ergonomics 26(10) (1983) p.963.
[122] W. Lotens, Heat exchange through clothing, in SafeWork Bookshelf, Jean-Jacques Vogt, ed.,
International Labour Organization, Geneva, 2006, Ch. 42.
[123] V. Yankelevich, Tekhnol. Legkoi Prom. 1 (1972) p.88.
[124] R. Nielsen, B.W. Olesen and P.O. Fanger, Ergonomics 28(12) (1985) p.1617.
[125] W. Lotens, Scand. J. Work Environ. Health 15 (1989) p.66.
[126] G.F. Fonseca, Text. Res. J. 45(1) (1975) p.30.
[127] M. Oguro, E. Arens, R. DeDear, H. Zhang and T. Katayama, J. Archit. Plann. Environ. Eng.
(561) (2002) p.21.
[128] Z. Vokac, V. Kopke and P. Keul, Text. Res. J. 43(8) (1973) p.474.
[129] P. Fanger, Thermal Comfort: Analysis and Applications in Environmental Engineering, Danish Technical Press, Copenhagen, 1970.
[130] M. Takahashi-Nishimura, S.S. Tanabe, Y. Hasebe and M.T. Nishimura, J. Physiol. Anthrop.
16 (1997) p.181.
[131] H. Belding, H. Russell, R. Darling and G. Folk, Am. J. Physiol. 149 (1947) p.223.
[132] R.R. Birnbaum and G.W. Crockford, Appl. Ergon. 9(4) (1978) p.194.
[133] J. Spencer-Smith, Cloth. Res. J. 5 (1977) p.3.
[134] B. Cain and B. Farnworth, J. Build. Phys. 9 (1986) p.301.
[135] J. Morris, Ann. Occup. Hyg. 17 (1975) p.279.
[136] Y.S. Chen, J. Fan, X. Qian and W. Zhang, Text. Res. J. 74(8) (2004) p.742.
[137] J. Fan and X. Qian, Eur. J. Appl. Physiol. 92 (2004) p.641.
[138] G.F. Fonseca and J.R. Breckenridge, Text. Res. J. 35(2) (1965) p.95.
[139] H. Anttonen and E. Hiltunen, The effect of wind on thermal insulation of military clothing, in
RTO Human Factors and Medicine Panel Symposium, NATO Science and Technology Organization, Neuilly-sur-Seine, 2009, p.1.
[140] G. Havenith, Technical evaluation report, in Blowing Hot and Cold: Protecting Against Climatic Extremes, NATO Science and Technology Organization, 2002, p.T1.
[141] D.McK. Kerslake, The Stress of Hot Environments, Cambridge University Press, Cambridge,
1972.
[142] C.D. Niven, Text. Res. J. 27(10) (1957) p.808.
[143] C.D. Niven, Text. Res. J. 29(10) (1959) p.826.
[144] R.J. Kind, J.M. Jenkins and F. Seddigh, Cold Reg. Sci. & Technol. 20(1) (1991) p.39.
[145] K. Cena and J.A. Clark, Phys. Med. Biol. 23(4) (1978) p.565.
[146] G.F. Fonseca and J.R. Breckenridge, Text. Res. J. 35(3) (1965) p.221.
[147] R.J. Kind and C.A. Broughton, Text. Res. J. 70(2) (2000) p.171.
[148] G.E.R. Lamb and M. Yoneda, Text. Res. J. 60(7) (1990) p.378.
[149] I.M. Stuart and E.F. Denby, Text. Res. J. 53(11) (1983) p.655.
[150] S.H. Lumley, D.L. Story and N.T. Thomas, Appl. Ergon. 22(6) (1991) p.390.
[151] E. den Hartog, Effects of clothing design on ventilation and evaporation of sweat, in Ergonomics of Protective Clothing (Proceedings of NOKOBETEF 6 and 1st European Conference
on Protective Clothing), Ingvar Holmer and Kalev Kuklane, eds., Arbetslivsinstitutet, Solna,
Sweden, 2000, p.281.
[152] H. Ueda and G. Havenith, The effect of fabric air permeability on clothing ventilation, in
Environmental Ergonomics, Yutaka Tochihara and Tadakatsu Ohnaka, eds., Elsevier
Science, 2005, p.343.
[153] W. Lotens and G. Havenith, Ventilation of rainwear determined by a trace gas method, in
Environmental Ergonomics, I. Mekjavic, E. Banister and J. Morrison, eds., Taylor & Francis,
Basingstoke, 1988, p.162.
[154] U. Danielsson, Convection cooling from wind and body motion, in Problems with Cold Work,
Ingvar Holmer and Kalev Kuklane, eds., Arbetslivsinstitutet, Solna, Sweden, 1997, p.260.
[155] Y. Satsumoto, J. Human Environ. Eng. 2(1) (2000) p.10.
[156] K.L. Harter, S.M. Spivak, K. Yeh and T.L. Vigo, Text. Res. J. 51(5) (1981) p.345.
[157] L. Berglund and T. Endrusick, Clothing ventilation estimates from manikin measurements, in
Sixth International Thermal Manikin and Modelling Meeting, J. Fan, ed., The Hong Kong
Polytechnic University, Hung Hom, Kowloon, Hong Kong 2006, p.158.
[158] N. Ghaddar and K. Ghali, Designing for ventilation in cold weather apparel, in Textiles for
Cold Weather Apparel, J.T. Williams, ed., Woodhead, Cambridge, 2009, p.131.
[159] K. Ghali, N. Ghaddar and E. Jaroudi, J. Heat Trans. 128(9) (2006) p.908.

TTPR_A_845540.3d (TTPR)

14-02-2014 17:22

Textile Progress
[160]
[161]
[162]
[163]
[164]
[165]
[166]
[167]
[168]
[169]

[170]
[171]
[172]
[173]
[174]
[175]
[176]
[177]
[178]
[179]
[180]
[181]
[182]
[183]
[184]
[185]
[186]
[187]
[188]
[189]
[190]
[191]
[192]
[193]
[194]
[195]
[196]
[197]
[198]

179

K. Ghali, M. Othmani, B. Jreije and N. Ghaddar, Text. Res. J. 79(11) (2009) p.1043.
N. Ghaddar, Int. J. Heat Mass Trans. 48 (2005) p.3151.
N. Ghaddar, Int. J. Therm. Sci. 42(6) (2003) p.605.
G.F. Fonseca and A.H. Woodcock, Text. Res. J. 35(10) (1965) p.909.
J.E. Ruckman, R. Murray and H. Choi, Int. J. Cloth. Sci. Technol. 11(1) (1998) p.37.
Y. Satsumoto, Y. Itou, Y. Hasebe and M. Takeuchi, Seni Gakkaishi 59(1) (2003) p.56.
Y. Satsumoto, W. Haihua, Y. Hasebe, K. Ishikawa and M. Takeuchi, Seni Gakkaishi 56(11)
(2000) p.524.
L. Fourt and M. Harrist, Text. Res. J. 17(5) (1947) p.256.
P. Gibson, Text. Res. J. 63(12) (1993) p.749.
W. Dyck, J. Cain and C. Moses, Determination of the water vapour resistance and thermal
resistance of sample materials using a sweating hot plate, in Defence and Civil Institute of
Environmental Medicine, Ontario, No. 96-R-65, Defence Research Establishment, Ottawa,
1997.
ISO 11092:1993, Textiles Physiological effects Measurement of thermal and resistance
under steady-state conditions (sweating guarded-hotplate test), 1993. Available at: http://
www.iso.org/iso/home/store/catalogue_tc/catalogue_detail.htm?csnumber=19081
K. Parsons, G. Havenith, I. Holmer, H. Nilsson and J. Malchaire, Ann. Occup. Hyg. 43(5)
1999, p.339.
R.L. Barker, Text. Res. J. 57(3) (1987) p.123.
R.M. Perkins, Text. Res. J. 49(4) (1979) p.202.
A. Spink, Waterproof/Breathable Garment Construction, US Patent 6308304, 2001.
S.J. Davies and M.B. Ducharme, Field Evaluation of Salomon Alpine Ski Clothing: Three
Multilayer Clothing Ensembles, Defence Research and Development Canada, Ottawa, 2001.
V. Krel, G. Hoffman, P. Offerman and K. Machova, Melliand Int. 5 (2005) p.E73.
Helly Hansen, H2 FlowTM Midlayer, 2012. Available at http://www.hellyhansen.com/press/
winter-12-13/h2-flow-midlayer.
B. Farnworth and R.J. Osczevski, Heat transport in cold weather clothing, in Prepared for
the Fourteenth Commonwealth Defence Conference on Operational Clothing and Combat
Equipment, Defence Research Establishment, Ottawa, 1985.
J. Gao, W. Yu and N. Pan, Text. Res. J. 77(8) (2007) p.617.
J. Gao, N. Pan and W. Yu, J. Text. Inst. 101(3) (2010) p.253.
B. Farnworth, Text. Res. J. 53(12) (1983) p.717.
IDFL, Brief Explanation of Down & Feather Tests, 2010. Available at http://www.idfl.com/
pdfs/IDFL Info - Brief Explanation of Tests/.
PHD-designs, Testing Down. Available at http://www.phdesigns.co.uk/techdown4.php?
[Accessed: 24-Jul-2013].
J. Gao and N. Pan, Text. Res. J. 79(12) (2009) p.1142.
J. Gao, N. Pan and W. Yu, J. Text. Inst. 100(6) (2009) p.539.
N. Du, J. Fan, H. Wu, S. Chen and Y. Liu, J. Theor. Biol. 248(4) (2007) p.727.
X. Wan, J. Fan and H. Wu, Polym. Test 28(7) (2009) p.673.
P. Gibson, D. Ph, C. Lee, F. Ko and D. Reneker, J. Eng. Fiber Fabr. 2(2) (2007) p.32.
J. Breckenridge, Insulating effectiveness of metallized reflective layers in cold weather clothing systems, in US Army Research Institute of Environmental Medicine, United States Army
Medical Research and Development Command, 1978, p.1.
D. Hegemann, M. Amberg, A. Ritter and M. Heuberger, Mater. Technol.: Adv. Perform.
Mater. 24(1) (2009) p.41.
X. Wan and J. Fan, Int. J. Heat Mass Trans. 55(2526) (2012) p.8032.
D.A. Holmes, Waterproof breathable fabrics, in Handbook of Technical Textiles, A. R.
Horrocks and S. C. Anand, eds., Woodhead, Cambridge, 2000, p.282.
G.R. Lomax, J. Ind. Text. 15(1) (1985) p.40.
J.C. Gretton, Moisture Vapour Transmission Through Outdoor Clothing Systems, University
of Leeds, Leeds, 1998.
J. Keighley, J. Ind. Text. 15(2) (1985) p.89.
D.W. Holden, J. Ind. Text. 17(1) (1987) p.46.
C.M. Carr, Chemistry of the Textiles Industry, 1st ed., Blackie Academic & Professional,
London, 1995.
R. Gore and S. Allen, Waterproof Laminate, US Patent 4194041, 1980.

TTPR_A_845540.3d (TTPR)

180

14-02-2014 17:22

M.P. Morrissey and R.M. Rossi

[199] R.J. Klare and D.E. Chubin, Porous Membrane Structure and Methods, US Patent 6410084,
2002.
[200] J.-C. Park, Electric Spinning Apparatus for Mass-production of Nanofiber, US Patent
7980838, 2011.
[201] B. Yoon and S. Lee, Fibers Polym. 12(1) (2011) p.57.
[202] S. Lee and S.K. Obendorf, Text. Res. J. 77(9) (2007) p.696.
[203] G. Havenith, E. den Hartog and S. Martini, Ergonomics 54(5) (2011) p.497.
[204] S. Lee and S.K. Obendorf, Fibers Polym. 8(5) (2007) p.501.
[205] P. Gibson, H. Schreuder-Gibson and D. Rivin, Colloids Surf. A 187188 (2001) p.469.
[206] H.W. Ahn, C.H. Park and S.E. Chung, Text. Res. J. 81(14) (2011) p.1438.
[207] W. Rietveld, Field Testing Air Permeable Waterproof-Breathable Fabric Technologies Part 2:
Are There Detectable Differences Under Real World Backpacking Conditions? 2011.
Available at http://www.backpackinglight.com/cgi-bin/backpackinglight/waterproof_breathable_
technologies_part2#.Ue-NGaI3BzM. (accessed: 24 July 2013).
[208] W. Rietveld, Field Testing Air Permeable Waterproof-Breathable Fabric Technologies Part
3: Discussion, Conclusions, and Performance of Individual Jackets, 2011. Available at http://
www.backpackinglight.com/cgi-bin/backpackinglight/
waterproof_breathable_technologies_part3. (Accessed: 24 July 2013).
[209] J. Holker, R. Jeffries and G.R. Lomax, Breathable, non-porous polyurethane film prepared
from a low molecular weight difunctional compound. US Patent 4367327, 1981.
[210] G.R. Lomax, Intelligent polyurethanes for interactive clothing, in 11th International Techtextil Symposium, Frankfurt am Main, 2001.
[211] A. Mukhopadhyay, J. Ind. Text. 37(3) (2008) p.225.
[212] R.J. Osczevski, Text. Res. J. 66(1) (1997) p.24.
[213] V. Bartels and H. Umbach, Text. Res. J. 72(10) (2002) p.899.
[214] P. Gibson, Polym. Test 19(6) (2000) p.673.
[215] J.C. Gretton, D.B. Brook and S.C. Harlock, Moisture vapour transmission through waterproof breathable fabrics under conditions of rain, in The Science of Climbing and Mountaineering: Proceedings of the 1st International Conference, University of Leeds, Leeds, 1999.
Neil Messenger, Will Patterson and Dave Brook, eds., CD-ROM, Human Kinetics
Publishers.
[216] R. Rossi, R. Gross and H. May, Text. Res. J. 74(1) (2004) p.1.
[217] Y.J. Ren and J.E. Ruckman, Int. J. Cloth. Sci. Technol. 16(3) (2004) p.335.
[218] E.a. McCullough, M. Kwon and H. Shim, Meas. Sci. Technol. 14(8) (2003) p.1402.
[219] K. Ledward, Condensation in Garment Systems KLETS. Available at http://www.klets.co.uk/
condensation_report.pdf. (Accessed: 24 July 2013).
[220] J.E. Ruckman, Int. J. Cloth. Sci. Technol. 9(1) (1997) p.23.
[221] J.E. Ruckman, Int. J. Cloth. Sci. Technol. 9(2) (1997) p.141.
[222] J. Zimmermann, F.A. Reifler, G. Fortunato, L.-C. Gerhardt and S. Seeger, Adv. Funct. Mater.
18(22) (2008) p.3662.
[223] J. Zimmermann, S. Seeger and F.A. Reifler, Text. Res. J. 79(17) (2009) p.1565.
[224] W. Barthlott, C. Neinhuis and C.L. Schott, Angew. Bot. 202 (1997) p.1.
 Ensikat and W. Barthlott, J. Exp. Bot. 55(397) (2004) p.711.
[225] E.
[226] S. Yoo and E. Kim, Text. Res. J. 78(3) (2008) p.189.
[227] C. Keiser, C. Becker and R. Rossi, Text. Res. J. 78(7) (2008) p.604.
[228] Q. Zhuang, S.C. Harlock and D.B. Brook, Text. Res. J. 72(8) (2002) p.727.
[229] Y. Wang, Z. Huang, Y. Lu, M. Zhao and J. Li, J. Text. Inst.104 (2012) p.1.
[230] N.P. Gao, H. Zhang and J.L. Niu, Indoor Built Environ. 16(1) (2007) p.7.
[231] A. Psikuta, D. Fiala, G. Laschewski, G. Jendritzky, M. Richards, K. Ba_zejczyk, I. Mekjavic,
H. Rintamaki, R. de Dear and G. Havenith, Int. J. Biometeorol. 56(3) (2012) p.443.
[232] J. Barry, R. Hill, P. Brasser, M. Sobera, C. Kleijn and P. Gibson, MRS Bull. 28 (2003) p.568.
[233] V. Bartels, Physiological comfort of biofunctional textiles, in Biofunctional Textiles and the
Skin, U.-C. Hipler and P. Elsner, eds., S. Karger AG, Basel, 2006, p.51.
[234] T.L. Vigo, J. Text. Inst. 90(3) (1999) p.1.
[235] S. Collie, Intelligent textiles are we overlooking the basics? Smarter routes to developing
smart textiles, in Technical Textiles: The Innovative Approach, Weston Conference Center,
UMIST, Manchester, UK, 2004.

TTPR_A_845540.3d (TTPR)

14-02-2014 17:22

Textile Progress

181

[236] M. Cybula, L. Rambausek, L. Van Langenhove and I. Krucin, Mater. Technol.: Adv. Perform.
Mater. 25(2) (2010) p.93.
[237] C. Hewitt, A.B. Kaiser, S. Roth, M. Craps, R. Czerw and D.L. Carroll, Nano Lett. 12(3)
(2012) p.1307.
[238] R. Rossi and D. Crespy, Polym. Int. 56 (2007) p.1461.
[239] A. Lendlein, Materials Today 11(3) (2008) p.59.
[240] B. DeCristofano, S. Fossey, E. Welsh, J. Perry and D. Archambault, MRS Proc. 1312 (2011)
p.137.
[241] K.J. Son, Impact Dynamics of Magnetorheological Fluid Saturated Kevlar and Magnetostrictive Composite Coated Kevlar, The University of Texas at Austin, Austin, 2009.
[242] L. Van der Schueren and K. de Clerck, Adv. Sci. Technol. 80 (2012) p.47.
[243] J. Hu, H. Meng, G. Li and S.I. Ibekwe, Smart Mater. Struct. 21(5) (2012) p.053001.
[244] Amazing Self-healing Revolution Bag. Available at http://www.edinburghbicycle.com/blog/
2012/05/healing-revolution-bag:/ (Accessed: 24 July 2013).
[245] L.E. Dorman and G. Havenith, Eur. J. Appl. Physiol. 105(3) (2009) p.463.

S-ar putea să vă placă și