Вы находитесь на странице: 1из 28

Journal ofhttp://jit.sagepub.

com/
Industrial Textiles

The Design of Waterproof, Water Vapour-Permeable Fabrics


G.R. Lomax
Journal of Industrial Textiles 1985 15: 40
DOI: 10.1177/152808378501500105

The online version of this article can be found at:


http://jit.sagepub.com/content/15/1/40

Published by:

http://www.sagepublications.com

Additional services and information for Journal of Industrial Textiles can be found at:

Email Alerts: http://jit.sagepub.com/cgi/alerts

Subscriptions: http://jit.sagepub.com/subscriptions

Reprints: http://www.sagepub.com/journalsReprints.nav

Permissions: http://www.sagepub.com/journalsPermissions.nav

Citations: http://jit.sagepub.com/content/15/1/40.refs.html

>> Version of Record - Jul 1, 1985

What is This?

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014


The Design of Waterproof, Water
Vapour-Permeable Fabrics
G. R. LOMAX
Shirley Institute
Didsbury, Manchester, U.K.

ABSTRACT
This paper reviews some current developments in fabrics which are both water-
proof and water vapour-permeable. Outer garments manufactured from these
materials can improve wearer comfort by reducing the buildup of perspiration in-
side the clothing. Tightly-woven fabrics and microporous polymer membranes
transmit water vapour predominantly by a diffusion-controlled mechanism similar
to air permeability. Apparently solid (i.e. non-microporous) polymer films and
fabric coatings which have much lower air permeability can also be designed with
good water vapour permeability. The hydrophilic mechanism involved is a com-
bination of (1) a physical process involving permanent or transient pores in the
molecular structure, and (2) an absorption-diffusion-desorption process which
depends on the chemical composition of the polymer and is specific for water
vapour.

INTRODUCTION
ability of outerwear fabric to transmit water vapour emitted from
an
The
the body as insensible perspiration and evaporated sweat is an impor-
tant factor in assessing the comfort of clothing assemblies [1]. Whilst a
modest buildup of water vapour can be tolerated with some discomfort,
the consequences of wearing relatively impermeable outer clothing may
be more serious in extreme weather conditions. In hot climates the wearer
may suffer from heat exposure. More commonly perhaps, accumulated
water vapour may condense inside the inner clothing of a wearer under-
taking arduous exercise in a cold, wet climate. Perspiration-soaked
clothing loses much of its insulative value and the wearer thus bears an
increased risk of body chill, or even hypothermia, when he or she ceases
activity. This condensation problem is particularly acute in the popular

40

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014


41

garments manufactured from continuously coated fabrics because the


standard polymers used (PVC, Neoprene, natural and synthetic rubbers,
and to a lesser extentacrylics, polyurethanes, and silicones) all have low
water vapour permeability.
A considerable amount of research has therefore been carried out on
elucidating the fundamental processes of water vapour transmission
through individual fabrics and clothing assemblies. The conclusions
drawn have proved extremely useful in the design of new outerwear
materials which combine maximum resistance to wind-driven rain and
snow with good water vapour transmission properties. Some of the

transport mechanisms involved, and the varied and interesting ap-


proaches taken to improve the comfort of foul-weather clothing are
described in the following sections.

TIGHTLY-WOVEN FABRICS
Water vapour and liquid water are transmitted through uncoated tex-
tiles by the following mechanisms:

(1) Simple diffusion through the interyarn spaces. This process is con-
trolled by the water vapour pressure gradient across the inner and
outer faces of the fabric. The resistance to diffusion is governed by
the fabric construction, i.e. the size and concentration of interyarn
pores and the fabric thickness.
(2) Capillary transfer through fibre bundles. In this mechanism liquid
water is &dquo;wicked&dquo; through the yarns and desorped or evaporated at
the outer surface. The efficiency of yarn wicking depends on the sur-
face tension, i.e. wettability of the fibre surfaces, and on the size,
volume and number of capillary spaces within the fibre bundle. The
nature of these interfibre spaces is determined by the choice of yarn
and fabric construction.
(3) Diffusion through individual fibres. This mechanism involves absorp-
tion of water vapour into the fibres at the inner surface of the fabric,
diffusion through the fibre structure, and desorption at the outer sur-
face. The ability of fibres to undergo water vapour diffusion depends
on the hydrophilic or hydrophobic nature of the fibre, Table 1 [2,3].
In open-weave fabrics, water vapour transmission occurs mainly
through interyarn spaces and transfer through individual fibres and fibre
bundles is relatively unimportant. Thus, fabrics of similar open construc-
tion, weight, and thickness are expected to show similar transmission
rates, irrespective of the type of yarn or fibre used. As the size of the inter-
yarn spaces decreases, however, these secondary transmission mech-

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014


42

Table 1. Moisture regain and water retention of some textile polymers.

Polyamide(lylon 11), polypropylene, PVC, PVDC, and glass, moisture regain (25°C, 65% r h ) less than
1.0, water retention less than 10 0 [3]
’Absorption of water vapour by dry fibres at 30°C and stated r h , m g water per 100g dry polymer [2].
2Water retained after centnfugation (0.25 h, 1000-1200 G) of soaked fibres, in g water per 100g dry poly-
mer [3]

anisms become moreimportant. Tightly-woven fabrics constructed from


absorptive or hydrophilic fibres such as wool, cotton, and viscose are con-
sequently more transmissive of water vapour than similar constructions
of non-absorptive, hydrophobic fibres [4]. A further advantage of natural-
fibre textiles is that they can quickly absorb substantial quantities of li-
quid water produced inside the clothing as sweat or condensed water
vapour. If a wearer is undertaking short, intermittent bursts of activity, for
instance, these materials can act as a comfort buffer by absorbing the
sweat produced directly from the skin and transmitting it to the ambient
atmosphere at a constant, diffusion-controlled rate.
Several authors have compared water vapour transmission through
woven fabrics with theoretical models based on vapour flow through the
uniform holes of a perforated metal plate. By varying the thickness of the
plate T, the fractional area of perforation B, and the diameter of each per-
foration d, the resistance to water vapour diffusion R was found to be
given by the expression [5]:

A fair to reasonable correlation was obtained between experimental


measurements on open weave fabrics and values of water vapour resis-
tance calculated from the dimensions of their interyarn pores using Equa-

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014


43

tion 1. These results stress the importance of fabric thickness and porosi-
ty to water vapour transmission through woven fabrics. This plate model
is of limited importance for certain tightly-woven fabrics, however, since
it assumes that transmission through the solid areas of the plate, cor-
responding to the contribution of fabric yarns and fibres, is zero-this is
clearly not the case for absorptive and hydrophilic fibres. If the metal
plate model is also applied to assess waterproofness, the water pressure P
required to penetrate the interyarn pores of a woven fabric can be approx-
imated from Laplace’s equation for the penetration of a liquid into a tube,
i.e.:

where r is the radius of the pore, y is the surface tension of water (or other
penetrating liquid), and 0 is the contact angle that the liquid makes with
the pore walls [6]. From Equation 2, the waterproofness of the fabric is
predicted to be increased by reducing the size of the interyarn pores, or
by increasing the contact angle through the use of a water-repellent
finish. Particularly high contact angles are obtained from hydrocarbons
such as purified paraffin wax (114°) and hexamethylethane (115 0) which
develop a smooth interface consisting mainly of strongly hydrophobic
methyl groups [7]. These materials are more effective water repellents
than hydrocarbons which contain only pendant hydrogen atoms, e.g.
polyethylene and various cyclic hydrocarbons, which have contact angles
of 105or less. In spite of their excellent water repellency, paraffin wax
and related hydrocarbons are unsuitable for use as permanent fabric
finishes because of their poor adhesion to the yarn and fibre surfaces.
Many commercial finishes are therefore based on amphipathic molecules
which consist of a highly polar, hydrophilic group attached to a long
hydrocarbon chain. When applied to fabrics, the polar ends of the
molecules become embedded in, or chemically bound to, the surface of
the yarns and fibres, and the hydrocarbon chains then tend to align with
their terminal methyl groups pointing outwards. If the inner hydrophilic
end of the molecule is sufficiently shielded, the surface of the finish
resembles the paraffin wax structure and generates high contact angles
(approximately 100-105°). This shielding effect increases as the number
of carbon atoms in the hydrocarbon chains increases, and reaches a max-
imum at sixteen carbon atoms and above. In addition to hydrocarbon-
based finishes, good water repellency is conferred by treating fabrics with
certain fluorocarbon and silicone systems. Because of their lower surface
tensions, perfluorocarbons tend to give slightly higher contact angles
with water than the corresponding hydrocarbons, and have the additional

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014


44

advantage of good oil repellency. The extreme hydrophobicity of the


perfluoromethyl and perfluoromethylene groups is demonstrated by the
fact that commercial fluorocarbon finishes, utilising a perfluorocarbon
chain containing only eight carbon atoms, have equivalent water
repellency to hydrocarbon-based finishes with chains containing sixteen
to eighteen carbon atoms. A recurring problem with chemically-finished
woven fabrics is deterioration in water-repellent properties over a period
of time, and garments may therefore require occasional reproofing. This
problem can be caused by surface soiling, particularly with any form of
detergent or surfactant, or the tendency of the contact angle to decrease
on prolonged exposure to liquid water. Some silicone finishes appear to
be less prone to a gradual reduction in contact angle, and are also less
susceptible to laundering than hydrocarbon-type finishes.
To summarise, the surface area and concentration of interyarn spaces
should be as high as possible to maximise water vapour transmission
through woven fabrics, and the fabrics should preferably be constructed
from absorptive and hydrophilic yarns. For foul weather clothing,
however, there is the contradictory requirement that the interyarn spaces
should be as small as possible to give maximum protection against wind
and rain, whilst the fabric outer surface should be non-absorptive and
hydrophobic to minimise wetting out by rain, sleet, or snow. Inevitably,
therefore, the design of the outerwear fabric must compromise between
waterproofness, windproofness, and water vapour transmission prop-
erties.
The first textiles with deliberately engineered waterproof and water
vapour-permeable properties were developed at the Shirley Institute just
before and during the Second World War. These all-cotton Ventile fabrics
are currently available in qualities ranging from the &dquo;stormproof&dquo; L19

grade (245 g/m2) used for outdoor occupations and pursuits, through to
the lightweight L34 grade (170 g/m2) used, for example, for surgeons’
gowns. The best all-round performance is given by the original L28 grade
(295 g/m2) developed specifically to resist penetration by sea water, and
thus prolong the survival time of aircrew forced to ditch in the cold North
Atlantic and Arctic seas. Although this latter grade is perhaps too stiff
and heavy for recreational use, L28 Ventile immersion suits are still stan-
dard issue for certain military personnel, and this fact highlights the
technical merits of the original design.
Ventile fabrics are woven from low-twist cotton yarns in the Oxford pat-
tern, i.e. plain weave with the warp threads run in pairs. This particular
construction imparts a flat surface with good abrasion resistance, and
maximises the closeness of the weave without overstiffening the fabric.
The yarns are spun exclusively from long staple Egyptian cotton and are

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014


45

mercerised before weaving to improve their regularity and swelling


capacity. In the final stage of production, the tightly-woven fabrics are
treated with a carefully formulated water-repellent finish and shrunk.
In the dry state, Ventile garments have a very low but measurable
degree of air permeability, and are generally regarded as windproof. The
combined effect of the air-permeable interyarn spaces and the hydrophilic
fabric construction ensures a good level of water vapour permeability,
and these materials have set the standard for outerwear comfort over the
past forty years. When the fabric is initially wetted by rainfall, or if the
wearer of an immersion suit ditches in the sea, the cotton yarns swell

rapidly and constrict the interyarn pores from approximately 10 microns


down to 3-4 microns across. It is the reduced pore size, together with the
applied finish, which prevents further penetration by rain or sea water.
Ventile garments thus have a higher degree of waterproofness than con-
ventional &dquo;showerproof&dquo; fabrics treated with a water-repellent finish;
however, compared with continuously-coated materials, they give only
moderate hydrostatic head values and may be penetrated by unusually
heavy rain or if long exposure times are involved. The choice of chemical
finish is also vital to fabric performance since some absorption of water
into the cotton yarns must occur to promote the pore constriction mech-
anism. The original finishes based on stearamidomethylpyridinium
chloride (e.g. Velan PF, Cerol WB) were found to be more effective water-
proofing agents than various silicone, fluorocarbon and wax emulsion
finishes, even though these latter materials probably generate larger con-
tact angles [8]. Presumably, these alternative finishes are too water
repellent and retard the initial swelling of the yarns, thereby allowing in-
gress of water through relatively larger interyarn pores.
Similar high quality garments are manufactured from cotton and
polyester/cotton fabrics which have been coated or impregnated with
proprietary finishes based on metallic soaps, hydrocarbon oils and waxes.
This type of clothing is also comfortable to wear, probably because the
fluid nature of the impermeable finishes prevents them from completely
blocking the fabric pores.
The latest developments in this area of textile research are based on
new technology for spinning extremely fine polyester filament yarns (0.1-
0.3 denier). When used in both warp and weft directions, the yarns can be
woven into various constructions, including taffeta, twill, and Oxford, in
which the individual filaments are packed so tightly that the fabrics
develop hydrostatic head resistances of 100-150 cm without the use of a
water-repellent finish [9,10]. Such fabrics are also claimed to have good
water vapour permeability and superior drapeability compared with
coated and laminated textiles.

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014


46

These types of garments, which have good waterproof and water


vapour transmission properties, commonly referred to as
are
&dquo;breathable.&dquo; Unfortunately, this term may be misinterpreted because the
breathability of a fabric is still sometimes confused with its air permeabili-
ty, which is essentially a measure of wind resistance. Since water vapour
is an intimate constituent of air, all woven, knitted and nonwoven textiles
which are considered to be air permeable by standard fabric test methods
should also transmit water vapour satisfactorily; however, the reverse is
not necessarily true. The majority of water vapour-permeable coated
fabrics described below fail to register a reading on air-permeability ap-
paratus devised for textiles, and these results are borne out by their good
windproof properties. However, this can be shown to be a limitation of the
test procedure, since, for example, microporous coated fabrics will
generate high air permeability values on more sensitive instrumentation
used to evaluate solid (non-microporous) packaging films and medical
products.
MICROPORO(1S POLYMER COATINGS
AND LAMINATED FILMS
When an open fabric is laminated to, or coated with, a continuous
polymer membrane, diffusion of water vapour through interyarn spaces,
yarns and fibres is severely reduced, and the permeability of the polymer
barrier becomes the rate determining factor for dissipation of perspira-
tion. One approach to increasing the inherently low water vapour
permeability of solid polymer membranes has been the development of
microporous structures with surface pores of much smaller dimensions
than the interyarn spaces of woven fabrics. Microporous films and
coatings function because of the enormous difference in size between
water vapour molecules (approximately 0.0004 microns in diameter) and
rain droplets which usually exceed 100 microns in diameter. The surfaces
of the microporous barrier are interconnected by a vast network of holes
and passageways. The pores at the outer surface are usually less than 2
microns in diameter and capable of repelling even the finest rainfall
through surface tension effects. Air and water vapour can pass freely
through the tortuous pathways provided that a suitable concentration
gradient exists between the inner and outer surfaces of the membrane.
The rate of water vapour transmission through a perforated foam struc-
ture has been shown to be related to its porosity (% pore volume) B,
thickness T, and pore diameter d, by the equation:

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014


47

where A is a [11]. For a constant porosity and thickness, this


constant
equation predicts that water vapour transmission through a foam struc-
ture increases as the pore size decreases.
With the increased popularity of coated textiles for fashionwear, rain-
wear and footwear during the late 1960s and early 1970s, a considerable
number of scientific papers and patents were published describing
methods for introducing such controlled porosity into solid polymers
(e.g. [12]). The simulated shoe leathers based on microporous polyure-
thanes manufactured by coagulation or salt extraction techniques were
possibly the most technically successful products of this period [13].
Lightweight apparel-grade versions of these synthetic leathers were also
introduced intermittently, but many failed to make a lasting impression
on the heavy-duty rainwear market.
One of the most significant developments in waterproof clothing was
therefore the introduction of the Gore-Tex laminates in 1976 [14]. The
functional component of these laminates is a thin, microporous mem-
brane made from solid PTFE sheet by a novel drawing and annealing pro-
cess. In drawn form, the tensile strength of the polymer is increased
threefold and its highly regular structure consists of PTFE nodules inter-
connected by numerous fine fibrils. This membrane is extremely perme-
able to water vapour on account of its high pore density and pore volume
(over 80%), whereas the fineness of the surface pores and high contact
angle of the base polymer bestow excellent waterproof properties. A
special adhesive system is used to laminate the microporous membrane
to a wide range of woven or knitted fabrics. The adhesive is applied to the
membrane in a discontinuous pattern which retains a high proportion of
surface pores available for air and water vapour transmission. In order to
overcome the well-known &dquo;non-stick&dquo; properties of PTFE, the adhesive

droplets must penetrate into the micropore structure to achieve me-


chanical keying, and some reduction in the high water vapour permeabil-
ity of the membrane is thus inevitable. The versatility of the lamination
process has contributed to the considerable success of the Gore-Tex
products during the past decade, since prototype fabrics and garments
can be readily manufactured to meet new outlets and performance re-

quirements as they arise.


A number of new microporous coated and laminated fabrics have also
been developed over the past five years, particularly in Japan where there
is a vast home market and export capacity for synthetic-based leisure-
wear. Most of these lightweight fabrics are based on porous polyurethane
direct coatings or laminated films which have been produced by wet
coagulation [9,10,15,16]. Microporous aluminium layers are included in
some grades of fabric, and are claimed to improve the warmth of a gar-

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014


48

ment by reflecting body heat [10,17]. However, it is doubtful that me-


tallised laminates can conserve much warmth, because heat loss through
the clothing occurs mainly by conductive and convective processes, not
by radiation [1].
New technology for producing microporous coatings has recently been
made available to traditional coaters who do not possess, or cannot afford
to install, sophisticated lamination equipment or coagulation lines. In one
process, for example, solid coated fabrics can be rendered microporous
by electron bombardment, and the operation carried out on a commis-
sion finishing basis [18]. This technique involves feeding the coated fabric
(up to 1.5 m wide) between two electrodes generating high-voltage elec-
tron beams which &dquo;drill&dquo; through the coating without damaging the fabric
beneath. The manufacturers claim that the electrons can be focussed into
discrete beams producing typically one million pores/m2 and with the ap-
propriate diameter for good water repellency.
Several European manufacturers of coated fabrics are currently using
the Ucecoat 2000 polyurethane system which can be applied to woven
substrates on existing direct-coating machinery. The recommended for-
mulation contains approximately 15-20% solids in a solvent mixture
containing methyl ethyl ketone, toluene, and a substantial proportion of
admixed water [19]. This product is therefore quite different from conven-
tional direct-coating polyurethanes which are applied as viscous solutions
containing 35-50% solids, usually in methyl ethyl ketone or ethyl
acetate. Fabrics spread with the Ucecoat 2000 formulation develop a
microporous coating during the drying and curing cycle, presumably by a
phase separation process. In such a process, the lower boiling solvent
(methyl ethyl ketone) evaporates preferentially as the fabric passes
through the oven, thereby increasing the concentration of non-solvent
(water) in the coating. When the concentration of non-solvent reaches a
critical level, the polyurethane precipitates out in a highly porous form
and the remainder of the solvents and non-solvents evaporate from the
coating as the fabric continues its passage through the oven. In contrast
to the wet coagulation process for synthetic leather production, no im-
mersion and washing baths are required to produce microporous
coatings by phase separation.

SOLID POLYMER COATINGS AND LAMINATED FILMS


In addition to transmission mechanisms involving interyarn spaces and
micropores, extremely small molecules such as the constituents of air, in-
cluding water vapour, can pass through solid polymers by a complex
molecular diffusion process [20]. A polymer film or coating which shows

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014


49

Photo 1. Surface of microporous PU coating based on Ucecoat 2000 system


(magnIfIcatIon 5,500x).

no evidence of voids or microporous structure can therefore still be


designed with good water vapour permeability. Because fabrics coated
with solid polymer contain no surface pores, they are not susceptible to
surface contamination and &dquo;pumping through&dquo; effects observed with
some microporous structures, and are exceptionally impermeable to
wind and rain.
The sensitivity of solid polymers to liquid water and water vapour is ex-
tremely variable and depends on their physical and chemical composi-
tion. At the two extremes of this spectrum are materials completely solu-
ble in water (polyvinyl alcohol, polyethylene oxide, etc.) and polymeric
compounds which are totally impermeable to water vapour (including
glass, diamond, and metals). In between these extremes, there are poly-
mers which can absorb large amounts of liquid water without being
dissolved (hydrogels) and other materials, which are not so heavily

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014


50

swollen by liquid water but nevertheless have a degree of water vapour


permeability. Elastomeric films and fabric coatings fall into this latter
category, see Table 2. In order to understand how these solid polymers
can transmit water vapour, and at greatly differing rates, some discussion
of their molecular structure is perhaps relevant.
All polymers consist of a series of long, flexible chains held together in
a three-dimensional network by weak attractive forces which may also be
reinforced by stronger chemical bonds (cross-links). The serviceable
temperature range of most flexible polymer films and coatings lies well
above their glass transition temperature; the molecular chains are
therefore in constant thermal motion and will tend to kink and fold ran-
domly along their lengths. This chain movement prevents perfect close
packing of the molecules, and very tiny holes or intermolecular pores are
formed throughout the polymer structure. Because the polymer chains
are constantly moving the molecular pores are frequently transient in
nature, and their size, density and distribution varies for different
polymers. Intermolecular pores are several orders of magnitude smaller
than micropores and, because of their transiency, are invisible even by
scanning electron microscopy. The constituents of air (oxygen, nitrogen,
carbon dioxide, water vapour, etc.) are still small enough to pass through
this ever changing pore system, although the transmission rates are con-
siderably slower than through a permanent microporous structure. It is
thus not strictly true to say that conventional polymer coatings are im-
permeable-they simply cannot transmit water vapour quickly enough
for an active wearer to remain comfortable.
The variations in transmission rates observed for different polymers are
related to the ease and degree of molecular packing, i.e. to polymer den-
sity, and to chain mobility, which depends on the chain stiffness and
cohesive forces between adjacent molecules. For example, the introduc-
tion of double bonds into the polymer backbone restricts close packing,
and unsaturated rubbers such as natural rubber, Neoprene and polybuta-
diene tend to have higher air and water vapour transmission rates than
fully saturated polymers. Because of the lower rotational energy of the
silicon-oxygen bond compared with the carbon-carbon bond, saturated
silicone rubbers have exceptionally high chain mobility and produce cor-
respondingly high transmission rates. Molecular motion decreases below
the glass transition temperature, and the size and volume of the inter-
molecular pore system is severely reduced. The transmission rates of
flexible films and coatings are therefore expected to decrease with
decreasing temperature. Rigid, &dquo;crystalline&dquo; polymers such as nylon,
polyester and other materials, which are below their glass transition tem-
peratures under normal conditions, are relatively poor transmitters of air

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014


51

and water vapour. Increasing the cross-link density, e.g. by vulcanisation


of rubbers or irradiation of polyolefins, also decreases the chain mobility
and therefore lowers the transmission rates. Conversely, the inclusion of
plasticisers or certain particulate fillers opens up the intermolecular pore
system of the polymer and increases the rate of gas and vapour diffusion.
The efficiency of water vapour diffusion through intermolecular pores
can also be increased by creating along the polymer chains a succession
of chemical groups (e.g. amine, hydroxyl, carboxyl), which are capable of
forming reversible hydrogen bonds with water vapour molecules. These
groups will effectively act as &dquo;stepping stones,&dquo; allowing water vapour
molecules to pass through the polymer from an area of high concentra-
tion to an area of low concentration. This hydrophilic mechanism is
specific for water vapour and does not involve oxygen, nitrogen, and car-
bon dioxide molecules which are incapable of forming hydrogen bonds.
As a secondary effect, however, the water vapour initially absorbed by the
polymer acts as a plasticiser by opening up the intermolecular pores and
also increases the diffusion of non-hydrogen bonding species. In most
polymer films and fabric coatings, the chemical groups required to ac-
tivate this hydrophilic mechanism are either totally absent (e.g. PVC,
PTFE, polyolefins, most hydrocarbon rubbers) or else they are in insuffi-
cient numbers and/or spaced too far apart along the polymer chains
(some acrylics, polyurethanes, etc.) to be effective. Certain hydrophilic
polymers possess the required concentration of hydrogen-bonding
groups with the correct chain spacing, and can transmit water vapour at
rates comparable with some microporous polymer membranes, when
evaluated under the same relative humidity/temperature gradients.
In the absence of pinholes, cracks, and other defects, transmission of
gases and vapours through solid polymer films and coatings therefore oc-
curs by absorption of the gas or vapour at the surface of high concentra-
tion, diffusion through the film at a rate controlled by the concentration
gradient, and desorption at the surface of low concentration [21]. The
transmission rate under steady-state conditions is given by the relation-
ship,

where (p, - pz) is the partial pressure gradient between the two surfaces, 1
is the thickness of the polymer, D is the diffusion constant, and S is the
solubility coefficient. The diffusion constant is governed by the nature of
the intermolecular pore system and, as already discussed, is affected by
the cohesive forces between polymer chains, crystallinity, density, and

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014


52

Photo 2. Microporous PTFE membrane (magnificafiion 5,500x) ,

modifications such as crosslinking, plasticisation, and inclusion of fillers.


For a given polymer at a constant temperature, the diffusion constant
also tends to decrease as the diameter of the penetrating gas or vapour
molecules increases. The solubility coefficient depends largely on the
molecular attraction between the polymer chains and the penetrant,
although there is a tendency for less volatile penetrants to condense in-
side the polymer structure and give high values. For simple inert gases,
the solubility coefficients are low and a linear relationship is given be-
tween the logarithm of the solubility coefficient and the boiling point of
the penetrating gas. The absorption of water vapour does not follow this
ideal behaviour, however, and both diffusion constants and solubility
coefficients are pressure dependent, particularly at high relative
humidities. The solubility coefficients of hydrophilic polymers are also
strongly affected by the high degree of polymer-water vapour interaction.
The water vapour permeability coefficients given in Table 1 are simply

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014


53

the product of the diffusion constant and solubility coefficient, i.e.


P = DS, and be substituted directly into Equation 4. It should be
can
noted that the hydrophilicity of a polymer can not be gauged from
measurements of water vapour permeability coefficients alone. For in-
stance, the high permeability coefficients of silicone rubbers are due to
their high diffusion constants and not to hydrophilic behaviour; in con-
trast, the high permeability coefficients of cellulose and its derivatives are
mainly associated with high solubility coefficients characteristic of hydro-
philic materials.
The concept of using hydrophilic polymers to selectively transfer water
vapour and other small molecules, whilst retaining liquid water, large
molecules and particles, is well documented (e.g. [22]). These materials
have been used for packaging films and for important medical applica-
tions, such as reverse osmosis and dialysis membranes, and more recent-

Table 2. Permeability coefficients of some common polymers (3).

’Permeability coefficients for oxygen are approximately 4 times those for nitrogen, carbon dioxide are ap-
proximately 25 times those for nitrogen [20].
2Glass transition temperature except * indicating complex second order transitions.

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014


54

ly for wound dressings. Some of the latest water vapour-permeable


coated fabrics in fact use a non-porous polyurethane membrane (Bion II)
which was developed originally for medical dressings [23]. Unfortunately,
however, most highly hydrophilic polymers are totally unsuitable for use
as permanent fabric coatings. A number of hydrophilic materials are too
sensitive to liquid water and, if used as waterproof coatings, would either
dissolve completely or become so heavily swollen by rain that they would
suffer severe flex or abrasion damage. Other polymers with a moderate to
high degree of crystallinity, notably regenerated cellulose and substituted
celluloses, are too stiff and inflexible. Incorporation of plasticisers would
soften these latter materials, although the additives are liable to migrate
or be leached out of the fabric coating under normal use.
Since the rate of water vapour transmission through a solid polymer is
inversely proportional to its thickness, thin polymer coatings with good
physical properties at low application weights are preferred for the hydro-
philic approach to water vapour-permeable rainwear. A moderate degree
of water vapour permeability coupled with good durability is thus ob-
tained with lightweight (15-50 g/m2) coatings of silicone rubbers and
some of their blends with polyurethane and acrylic formulations [24]. The
use of polymer blends incorporating hydrophilic polymers or pigments,

e.g. mixtures of poly(ethyl acrylate) and polyvinyl alcohol, has also been
advocated [25]. Some aqueous-based and one-component polyurethanes
used in transfer coating are also considered to be water vapour-
permeable, particularly when they are applied to open-weave, knitted and
cellulosic fabrics.
A direct coating system with distinctly improved water vapour per-
meability has recently been introduced by Baxenden Chemicals [26].
Their Witcoflex 971 product is based on polyurethanes, developed
originally at the Shirley Institute [27], in which the hydrophilic com-
ponents are permanently bound into the molecular structure, and hence
cannot separate out or be leached out of the coatings in use. The poly-
urethane mixture is supplied as 45% solids dissolved in methyl ethyl
ketone, and can be cross-linked with standard additives, such as a pro-
prietary tri-isocyanate. Witcoflex 971 formulations thus resemble most
two-component polyurethanes in appearance and spreading rheology,
although some technical expertise is required to obtain the maximum
performance from a particular direct coating/fabric combination.
The construction of the base fabric and method of polymer application
also contribute significantly to water vapour transmission through solid
polymer membranes. Direct coating is normally carried out with tightly-
woven fabrics manufactured from nylon or polyester filament-spun yarns.
In order to achieve good mechanical adhesion of the coating to the

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014


55

Photo 3. Surface of Witcoflex 971 solid coating-note absence of pores and smooth
abraslon-resistant surface (magnificatIon 5,500x).

smooth fabric surface, the polymer spreading solution must penetrate


deep into the fabric interstices; excessive penetration or &dquo;strike-through&dquo;
must be avoided, however, as this imparts an undesirably stiff handle and
crackle to the coated material. Direct polymer coatings thus effectively
consist of a mesh of relatively thin sections, representing the areas of
polymer covering the yarn crowns, and thicker sections where it has
penetrated into the fabric interstices. Transmission of water vapour oc-
curs mainly through the interstices of non-absorptive, hydrophobic
fabrics, and hence through the predominantly thicker areas of the poly-
mer barrier. By comparison, cotton and polyester/cotton substrates can
also utilise the hydrophilic properties of the fibre bundles under the thin-
ner areas of polymer coating, and consequently tend to be more water

vapour permeable than coated nylon and polyester fabrics of similar


structure. The thickness of the polymer membrane is less variable in
transfer-coated and film-laminated textiles, and open-weave and knitted

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014


56

Photo 4. Section through Witcoflex 971 solid PU direct coated onto woven nylon
fobnc-note absence of pores and excellent mechanical adhesion to nylon filaments
(magnification 1,800x)

fabrics with larger surface areas available for water vapour transmission
are often used instead of tightly-woven substrates. For a given polymer

coating weight, therefore, transfer-coated and laminated fabrics are more


permeable than direct-coated fabrics of the same fibre type.

EVALUATION OF FABRICS
It is not intended to compare critically the performance of the various
fabrics and coating systems described above, but to discuss briefly some
salient methods for their evaluation. The main functional parameters for
foul-weather clothing designed primarily for temperate climates are pro-
tection against wind, rain and cold, durability, launderability, and com-
fort. With the introduction of new materials and increased competition,
the non-functional properties such as appearance, aesthetic handle,

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014


57

design, availability and, of course, cost become increasingly more impor-


tant.
At the present time, there is no single standard relating specifically to
all types of waterproof, water vapour-permeable fabrics and garments,
possibly because of their diversity of design and construction. Many of
the relevant parameters can be measured by standard test methods
relating to conventional waterproof fabrics, e.g. testing coated fabrics
(BS 3424:1982) and the specifications for lightweight coated fabrics
(BS 3546:1981) and foul-weather clothing (BS 6408:1983). These stan-
dards invariably omit methods for assessing comfort.
The waterproof properties of a fabric can be easily evaluated by a
number of different static and dynamic methods [6,14]. Most tests
relating to proofed, air-permeable fabrics measure either the resistance of
a fabric to surface wetting and absorption or its penetration by simulated

heavy rainfall; however, some methods evaluate both aspects of water


repellency simultaneously. In the Bundesmann test, for example, the
fabric is exposed to a simulated rain shower while its under surface is sub-
jected to a rubbing action, thus imitating body movement inside a gar-
ment. The amount of water absorbed by, and passing through, the fabric
is then measured after a period of time, usually ten minutes. It has been
argued that this test is a particularly severe form of rain simulation, since
the kinetic energy of the water droplets produced is much higher than
that produced in actual rainfall (Table 3), and may therefore reject fabrics
whose performance may be acceptable in normal use [28]. While the
Bundesmann test is still used by some manufacturers, mainly for quality
control and screening purposes, there appears to be too much variation
between results obtained with different pieces of equipment for the
method to be designated as a British Standard; consequently, the
Bundesmann method has been superseded by the similar WIRA Shower
test (BS 5066:1974) for many applications.

Table 3. Comparison of rain Intensity (28).

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014


58

Photo 5. Ventile fabric in the dry state-note the interyarn pores (magnification
unknown).

Highly waterproof materials are not usually penetrated by this type of


simulated rainfall, and so more discriminating tests for measuring water
entry pressure, e.g. hydrostatic head (BS 2823:1982), have been
developed. Coated fabrics, as received from the manufacturer, usually ex-
ceed 150 cm hydrostatic head which is the upper limit for some types of
apparatus; the actual entry pressure can be considerably higher, par-
ticularly for solid coated fabrics and PTFE laminates [14]. Proofed cotton
fabrics produce lower hydrostatic head values, e.g. approximately
130-140 cm for an L28 grade Ventile, and 100 cm or less for lighter
grades. Since hydrostatic penetration tests effectively measure the in-
tegrity of the polymer barrier, they can also be used to assess the dura-
bility of the polymer coating or laminated film under a wide range of wear
and accelerated ageing conditions. Most coated garments are designed
with the polymer barrier on the inside to protect it against external
snagging and abrasion damage. The outer fabric face is therefore routine-

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014


59

Photo 6. Ventile fabric after initIal wetting by raInfall-note that cotton yarns have
swollen and constricted interyarn pores

ly treated with a silicone or fluorocarbon finish after the coating or


laminating operation has been carried out. This treatment prevents the
garment from wetting out by pearling the raindrops off the outer surface,
and its effectiveness can be evaluated by various spray tests, e.g. BS
3702:1982, before and after laundering.
Windproofness can be evaluated by the air permeability test (BS 5636:
1978). Most coated fabrics have lower air permeability than Ventile
fabrics, which have proven windproof qualities, and can therefore be ex-
pected to be adequate in this respect.
The durability of uncoated woven fabrics can be evaluated by measure-
ments of tear strength, tensile strength, and abrasion and flex resistance.
Additional tests carried out on coated and laminated fabrics should in-
clude peel strength, abrasion and flex resistance to measure wet and dry
adhesion of the polymer coating to the base fabric. It has also been
recommended that the flex testing programme should be carried out at

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014


60

Photo 7. Section through Cyclone 3-layer laminate (magnification 7,800x)

5°C and progressively lower temperatures, and the specimens then


tested to hydrostatic failure [29]. These conditions are thought to
simulate more closely the temperature of wind-driven rain, and are a
more realistic and severe test for delamination resistance.

Launderability may be assessed by subjecting the fabrics to a series of


soiling and washing or drycleaning cycles, and testing the samples for
reduced hydrostatic head properties. Coated fabrics can also be tested for
reduced coating adhesion after laundering. Water-repellent finishes, oils,
and waxes may be rendered ineffective or removed completely by dry-
cleaning or washing in warm, soapy water. In order to avoid constant
reproofing, these finished fabrics should be washed in cold, detergent-
free water on a non-routine basis. Microporous coated fabrics are also
susceptible to loss in waterproof properties if they are not thoroughly
rinsed out after laundering. This problem arises when surfactants or

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014


61

Photo 8. A typical transfer coated fabric Solid polyurethane, knitted nylon fabric-note
large area of the film which is not attached to fabnc (magnIfIcatIon 780x)

detergents present in washing powders and commercial drycleaning


solvents are entrapped in the surface pores, and, by lowering the surface
tension of the rain droplets and their contact angle with the pore walls,
reduce the water penetration pressure according to Equation 2. Pho-
tographic evidence suggests that some manufacturers of microporous
coated fabrics are now sealing the surface pores of the barrier membrane
with a very thin, possibly hydrophilic, solid coating. This measure should
greatly reduce the problems of surface contamination, without signifi-
cantly lowering water vapour permeability.
The comfort aspect of foul-weather clothing, as of any apparel textile,
is extremely difficult to quantify, and a detailed review of the physiologi-
cal factors involved is beyond the scope of this paper. Some of these
aspects have been discussed previously in this journal in relation to the
comfort of Gore-Tex laminates [14,30,31]. The thermal resistance and

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014


62

water vapour resistance of a fabric are important comfort parameters


since they affect the heat loss through the clothing by conduction and
evaporation according to Woodcock’s equation,

where (T. - T.) and (ps - p.) are the differences in temperature and water
vapour pressure, respectively, between the skin and ambient environ-
ment, and I and E are the total thermal resistance and water vapour
resistance of the clothing and intervening air layers [32]. The measured
thermal and water vapour resistance of fabrics are directly additive i.e.,
clothing assemblies can be evaluated by summation of the individual
values of the fabric components and the air layers between them. The im-

Photo 9. Section through Cyclone fabric, Woven upper fabric, microporous


polyurethane mid-layer and lower knitted fabric laminated to PU with discontinuous
adhesive web (magnification 7BOx).

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014


63

portance of the air layers should be stressed, since they often constitute
up to 80% of the total resistance of the clothing assembly.
Thermal resistances of fabrics can be readily measured, e.g. by
BS 4745:1971, and lightweight proofed cotton and coated fabrics pro-
duce low values, i.e. they usually have little insulation value. If the
clothing is worn sensibly using a &dquo;layer&dquo; system, the inner layers provide
warmth and the outer layer is designed to offer maximum protection
against inclement weather. With this type of clothing assembly, it is
perhaps desirable that the thermal and water vapour resistances of the
outer layer be as low as possible, as this condition maximises the theo-
retical heat flow through the outer garment, and allows greater versatility
in the choice of inner clothing [33]. If required, however, the warmth of
the outer garments can be improved by incorporating a loose or bonded
lining having some insulative value, or by using the waterproof fabric as
the outer material in a duvet-style garment.
Water vapour resistance can also be measured by a number of different
methods [1], and should always be as low as possible for an outerwear
fabric, although this is inevitably limited by the degree of rain and wind
protection required. Whilst textile researchers and physiologists use
water vapour resistance values for assessing the ability of a fabric to
transfer perspiration, fabric and garment manufacturers often prefer to
know water vapour transmission rates expressed in more familiar units of
g/m2/day. Surprisingly however, the water vapour transmission rate of an
outerwear fabric is one of the most difficult and controversial properties
to measure. Considerable variation in the magnitude of transmission, and
also in the order of ranking of a set of fabrics, can be observed on chang-
ing from one of the numerous test methods available to another. These
discrepancies usually arise when fabrics are evaluated under quite dif-
ferent temperature and relative humidity gradients, and so care should be
taken when comparing transmission values obtained from different
sources.

Although thermal and water vapour resistances of fabrics are useful for
evaluating static clothing assemblies, they are not completely satisfac-
tory for the physiological assessment of garments in actual use. This
shortcoming is due to the fact that the air layers between the fabric com-
ponents which govern temperature and relative humidity gradients are
constantly changing during body movement, and the wearer is also sub-
ject to variations in body metabolism, intermittent sweating and con-
siderable changes in ambient temperature, relative humidity, and wind
velocity. The design of the garment is also important for comfort, espe-
cially features which allow the wearer to adjust ventilation. Clothing
physiologists therefore recommend a tiered system for evaluating the

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014


64

comfort of new clothing materials [1,34]. Typically, this method involves

laboratory measurements of fabrics and garment assemblies under static


conditions and a range of wear-simulated conditions [35], followed by
physiological trials carried out in a climatic chamber, and leading even-
tually to limited and long-term field trials.
Finally, it should be emphasised that the water vapour resistance of the
clothing is only one of a number of variables in the comfort equation;
climatic conditions which substantially reduce the water vapour pressure
gradient across a garment or clothing assembly also restrict evaporative
heat transfer. Moisture may therefore still accumulate inside a normally
comfortable outer garment worn, for instance, on a warm, humid day
with little or no wind present.

SUMMARY
There are a number of tightly-woven and coated fabrics currently avail-
able which can substantially reduce the buildup of perspiration inside
waterproof clothing. In addition to the civilian and military markets for
heavy-duty rainwear, these materials are of potential use in a number of
applications where accumulation of water vapour is undesirable. These
outlets may include immersion suits, protective workwear, footwear, in-
cluding panel inserts and shoe liners [36], speciality tarpaulins and fitted
covers, tentage, sleeping bag liners and mattress covers. Surgeons’
gowns are a particularly interesting example where the fabric must act as
a total barrier against bacteria, skin debris and body fluids.
In spite of the popularity of lightweight coated outerwear, Ventile-type,
oiled and waxed garments have an aesthetic, even fashionable, appear-
ance which is preferred for some pursuits e.g., hunting, fishing, and

shooting; these markets have only recently been challenged by the better
quality coated and laminated fabrics. Because of revived interest in the
&dquo;traditional&dquo; look, cotton and polyester/cotton blends are being increas-
ingly used as coating substrates, together with texturised nylon fabrics
(e.g. Taslan, Tactel) which are less harsh in appearance and handle than
conventional filament-woven fabrics.
Some of the high-performance products mentioned herein are justifi-
ably more expensive than normal waterproof fabrics, and some others are
not amenable to standard production methods. Thus, even though con-
siderable progress has been made during the past decade, there is still
ample scope for further technical and commercial developments in water
vapour-permeable foul-weather clothing.
ACKNOWLEDGEMENT
The author wishes to thank the Ministry of Defence and the Depart-

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014


65

ment of Trade and Industry for providing financial assistance towards


Shirley Institute projects on the development of water vapour-permeable
coatings.
REFERENCES
1. "Textiles for Comfort," Proceedings Third Shirley International Seminar,
Shirley Institute, Manchester, 1971.
2. Jeffries, R., J. Textile Institute, 51
, T339, T399, 1960.
3. Roff, W. J. and J. R. Scott, Fibres, Films, Plastics and Rubbers. Handbook of
Common Polymers, Butterworths, London, 1971.
4. Fourt, L. and M. Harris, Textile Research J., 17
, 256, 1947.
5. Whelan, M. E., L. E. MacHattie, A. C. Goodings, and L. H. Turl, Textile
Research J., 25, 197, 1955.
6. Waterproofing and Water-Repellency
, ed. Molliet, J. L., Elsevier, London,
1963.
7. Adam, N. K. and J. E. P. Elliot, J. Chem. Soc., 2206, 1962.
8. Holker, J. R. and G. Lund, Lenzinger Berichte, 43, 66, May 1977.
9. Anon., Japan Textile News, 32, January 1984.
10. Anon., Japan Textile News, 32, March 1983.
11. Fonseca, G. F., Textile Research J., 37, 1072, 1967.
12. Sittig, M., Synthetic Leather from Petroleum, Noyes Data Corporation, 153,
New Jersey, 1969.
13. Fukushima, O., J. Coated Fabrics, 5, 3, 1975.
14. Gohlke, D. J. and J. C. Tanner, J. Coated Fabrics, 6, 28, 1976.
15. Anon., Japan Textile News, 24, November 1983.
16. Anon., Japan Textile News, 32, December 1982.
17. Anon., Japan Textile News, 20, May 1983.
18. Anon., Melliand Textilberichte, 64, 493, 1983.
19. Ucecoat 2000 Data Sheet, UCB Chemical Sector, Drogenbos, Belgium.
20. Diffusion in Polymers, ed. Crank, J. and G. S. Park, Academic Press, New
York, 1968.
21. Rogers, C., J. A. Meyer, V. Stannett, and M. Szwarc, TAPPI, 39, 737, 1956.
22. Kesting, R. E., Synthetic Polymer Membranes
, McGraw-Hill, New York, 1971.
23. Mansfield, R. G., Textile World, 58, May 1985.
24. Clark, P. J., Polymer Surfaces
, ed. Clark, D. T. and W. J. Feast, Wiley-
Interscience, Chichester, 235, 1978.
25. Caldwell, J. R. and C. C. Dannelly, American Dyestuff Reporter, 56, P77, 1967.
26. Witcoflex 971 Data Sheet RF 23, Baxenden Chemical Co. Ltd., Droitwich,
England.
27. Anon., Textile Asia, 15, 130, August 1984.
28. Baxter, S. and A. B. D. Cassie, J. Textile Institute, 36, T67, 1945.
29. Keighley, J. H., Breathability in Outer Fabrics, presented in part at the Survival
’85 Conference, Leeds, March 1985.

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014


66

30. Tanner, J. C., J. Coated Fabrics, 8


, 312, 1979.
31. Gohlke, D. J., J. Coated Fabrics, 10
, 209, 1981.
32. Woodcock, A. H., Textile Research J., 32, 628, 1962.
33. Greenwood, K., W. H. Rees, and J. Lord, Textile Institute Diamond Jubilee Con-
ference, London, 197, 1970.
34. Meechels, J. H. and K. H. Umbach, Clothing Comfort, ed. Hollies, N. R. S. and
R. F. Goldman, Ann Arbour Science, Ann Arbour, 133, 1977.
35. Steinhaus, I. and G. Eissing, Melliand Textilberichte, 65
, 145, 1984.
36. Duncan, J., J. Coated Fabrics, 13, 161, 1984.

EDITOR’S NOTE
The technology described in this contribution by Dr. G. R. Lomax is be-
ing commercialized under a licensing agreement with Baxenden
Chemical Co., Ltd. Baxenden is now licensed to manufacture and sell
polyurethanes based on technology developed at Shirley Institute and
designed primarily for use as waterproof, water vapour-permeable (i.e.
breathable), non-poromeric coatings on textile and other substrate
materials.
The novel polymers were synthesised at Shirley Institute by Dr. J. R.
Holker and Dr. G. R. Lomax in the course of research work supported ini-
tially by the Ministry of Defence and subsequently, in part, by the Depart-
ment of Trade and Industry via the Garment and Allied Industries
Requirements Board*. The aim of this research has been to develop
breathable waterproof fabrics. Garments, such as jackets or anoraks,
made from those fabrics transmit water vapour (perspiration) away from
the body as rapidly as it is generated and greatly enhance wearer-comfort,
especially for outdoor pursuits or occupations where strenuous activity is
being undertaken under adverse weather conditions.
Baxenden Chemical Company has contributed further development
work on the formulation and application of the polymer and is now manu-
facturing a polyurethane coating solution based on the new technology.
In addition to its breathable properties, the polymer coating possesses
good physical properties, has a &dquo;soft&dquo; feel, and is easy to apply to suitable
textile materials.
Baxenden states that the new coating material, WITCOF’LEX 971, is
already being used by several of the UK’s outdoor waterproofs manufac-
turers in the production of anoraks and related garments and that there is
extensive worldwide interest in the innovative coating technology.

’This Board has now been superseded by TOMRB, the Textiles and Other Manufactures Re-
quirements Board.

Downloaded from jit.sagepub.com at St Petersburg State University on February 10, 2014

Вам также может понравиться