Está en la página 1de 9

Journal of the Human-Environmental System

Review Article Vol. 5; No. 2: 69–77, 2002

Air and Air Spaces—the Invisible Addition to Thermal Resistance

C. A. Wilson, R. M. Laing and D. J. Carr1)


1)
C. A. Wilson and D. J. Carr are Lecturers and R. M. Laing an Associate Professor in Clothing and
Textile Sciences at the University of Otago, Dunedin, New Zealand
E-mail: c.wilson@otago.ac.nz

(received on December 21, 2001, accepted on April 20, 2002)

Abstract
The aim of this review is to examine the contribution which air makes to the overall thermal resistance of
garments (primarily) when in use. Issues related to i) air contained in fabrics, between fabric layers, between
fabrics and the body, ii) garment design and the presence and effectiveness of closures and vents, and iii) the
effects of wind and body movement are addressed. How these issues are affected by changes to body posi-
tion, movement of air within the garment microclimate and exchange of the garment microclimate air with am-
bient air are also discussed. Indicators of a change in thermal resistance have generally been restricted to a
change in microclimate air movement or to a change in energy required to maintain a thermal plate/manikin
at a fixed temperature.

Key words: thermal resistance, air, textiles, layers, gaps, garment fit

Introduction ers of fabrics, between fabrics and the human body


During development of many types of garments, and movement of air within a garment microclimate
fabrics which are essentially two-dimensional are and exchange with ambient air; 2) issues related to
transformed into three-dimensional shapes. The ac- garment design such as posture of the wearer, gar-
tual measurements of the finished garments and those ment fit, and the presence and effectiveness of clo-
of the intended user are not only related to appear- sures and vents; and 3) the effects of wind and body
ance, but also to the thermal resistance of that gar- movement.
ment. The size and shape of air layers and their ef-
fects on thermal resistance vary depending on factors 1. Air
such as some physical characteristics of the fabrics, Air in textiles
configuration of the clothing e.g. posture and cover or All yarn and fabric arrangements consist largely of
wrapping arrangement, activity of the wearer, and fiber and air (ignoring most finishes), with fiber dom-
ambient environmental conditions. The aim of this inating in terms of mass and visibility and air domi-
review is to examine published research on the con- nating in terms of volume. For example, the volume
tribution which air makes to the overall thermal re- of air present in woven fabrics ranges from 60–90%
sistance of garments (primarily) when in use. In this and in knits from 85–95%, while products such as
review, indicators of a change in thermal resistance quilts are reportedly comprised of as much as 95–
have been restricted to evidence of a change in move- 99% air (Schneider and Holcombe, 1988). As the
ment of microclimate air or to a change in energy re- thermal conductivity of fibers ranges from 5–20
quired to maintain a thermal plate/manikin at a fixed times that of air (Holcombe, 1984; Holcombe, et al.,
temperature. Generally no attempt has been made to 1988), it is the air trapped within the yarn, fabric, and
review the extensive literature on changes in human garment (or product) itself that contributes most to
performance resulting from a change in thermal re- the overall thermal resistance.
sistance of a garment or garment assembly, as this The major contribution of still air to thermal resist-
was part of a comprehensive review on clothing, tex- ance of flat textiles has lead to thickness of materials
tiles, and human performance currently in press being used to estimate thermal resistance. However,
(Laing and Sleivert, 2002). thickness is only an appropriate predictor if a stan-
There are three main parts to the present review: 1) dard value of thermal resistance per unit thickness is
issues related to air contained in fabrics, between lay- appropriate, and thickness of materials (difficult to
70 C. A. Wilson, R. M. Laing and D. J. Carr

determine), is measured accurately. For example, layers in form-fitting clothing are considered to in-
many fabrics considered to be effective insulators, crease thermal resistance from 5 to 50% (increases
have a high proportion of projecting surface fibers are particularly apparent in relatively small air
which trap air and instruments commonly used to spaces) (Havenith, et al., 1990; Holcombe, 1984;
measure thickness cause compression of the speci- Morris, 1953; Morris, 1955; Weder, et al., 1996).
men. These factors affect the reliability of the thick- In infant bedding, where air spaces of 180 mm in
ness:thermal resistance relationship (Bogaty, et al., height have been reported, increases of up to 100% in
1953). The linearity of the relationship between ‘dry’ and 300% in ‘wet’ thermal resistance of dry
thickness and ‘dry’ thermal resistance was ques- fabrics have been attributed to the air layer (Wilson,
tioned as early as 1955 and in the case of multiple et al., 1999a; Wilson, et al., 2000). Using a trace gas
layers of thin, lightweight materials (and the air lay- method, the introduction of an air layer under a jacket
ers between them), has been shown to be non-linear (size and volume not specified) by the use of a spacer
under certain conditions (Epps and Song, 1992; Mor- showed the exchange of the jacket microclimate air
ris, 1955). with air from the ambient environment increased by
140–300% (Lotens and Havenith, 1988). Several
The air layer—effect of thickness of air layers on smaller air spaces have also been shown to provide
thermal resistance greater thermal resistance when distributed through-
Air layers may form under, between, and over fab- out the bedding than when the same total thickness
ric surfaces contributing to the thermal resistance of occurred as one large air space under the first layer
multiple-layer materials. The properties of these air (Wilson, et al., 1999a). Both the size and distribution
layers are affected by their thickness, shape, and dis- of air layers affect thermal resistance.
tribution. When two-dimensional fabrics are wrapped
around the human body the geometry of the surface Variables affecting distribution of air layers within
changes, while stacking fabrics one over another may fabrics and garments
result in additional air being enclosed between fab- Many investigators have attempted to define vari-
rics, and not necessarily uniformly distributed. Tex- ables which affect the distribution and thickness of
tile covers are typically non-homogeneous as the air spaces within, and thermal resistance of, a gar-
proximity of the fabric layers to each other and the ment assembly. Variables such as i) fabric character-
number of layers may vary over the surface of body. istics (including packing density, fabric mass, fiber
For example, the head and hands are often covered fineness, yarn count and twist, hairiness, compress-
only with stable still air (i.e. the boundary layer), ibility) (e.g. Holcombe, et al., 1983; Obendorf, et al.,
which contributes approximately 0.02–0.12 m2K/W 1986); ii) fabric manufacturing processes (such as
to the thermal resistance of the body (Goldman, surface napping and application of finishes); iii) style
1988; Holcombe, 1984). This boundary layer thus of garment (including fit, multiple layers) (e.g.
provides some thermal protection to both clothed and Lotens, 1989), fit and product design (e.g. Fan and
nude bodies (although the contribution of the bound- Keighley, 1989; McCullough, et al., 1983), and sur-
ary layer to thermal resistance is considerably re- face area and geometry effects (e.g. Spencer-Smith,
duced in moving air). 1977) and iv) conditions of use (such as the ambient
Thermal properties of fabrics with enclosed or un- environment, air movement; and intensity and type of
derlying air layers have been investigated using labo- activity of the wearer/user) are some of those that
ratory techniques, both flat (e.g. Cain and Farnworth, have been evaluated. Many of these issues/variables,
1986; Wilson, et al., 1999a; Wilson, et al., 2000; as they relate to the textile, have been addressed in
Woo, et al., 1994), those simulating the 3-dimen- the literature (Holcombe, 1984; Peirce and Rees,
sional arrangement of materials during use (e.g. Bo 1946; Ukponmwan, 1993). However, garment design
and Nakajima, 2001; Fonseca, et al., 1965), and in (McCullough, et al., 1983), the trapping of air, and
actual use (e.g. Lotens, 1989; Nielsen, et al., 1985). conditions of use (Havenith, 1999; Havenith, et al.,
As the size of the air gap increases the relationship 1990; Lotens, 1989) have become the focus of sys-
between thermal resistance and the number of layers tematic research only from the 1980s.
has been shown to become non-linear (Epps and
Song, 1992; Wilson, et al., 1999a; Wilson, et al., Convection within air spaces
2000). Wilson, et al., (1999a) showed the magnitude Convection involves the transmission of energy or
of this difference was affected by the thickness of the mass by movement of the medium and occurs as a re-
underlying air gap, the arrangement of the layers, the sult of the movement of large masses of molecules in
number of layers, and the distribution of air layers the form of air currents (Parker, 1994). Air move-
within the assembly. Air gaps between the component ment occurs as a result of natural or forced convec-
Air and Air Spaces—the Invisible Addition to Thermal Resistance 71

tion with the rate of convective heat loss affected by the air space increased. ‘Wet’ thermal resistance pro-
the air velocity and the temperature gradient. The ef- vided by an air space under a standard dry fabric in-
fect of convection in air spaces on thermal resistance creased to a point (40 mm) beyond which any fur-
has been investigated, generally using thermal hot- ther increases in the thickness of the air space did not
plates (e.g. Fowle, 1933; Hassenboehler and Vigo, increase thermal resistance (Wilson, et al., 2000).
1982; Hollands, et al., 1975; Hollands, et al., 1976), The orientation and thickness of the air layers ap-
humans (e.g. Ruckman, et al., 1999), and articulated pears to also affect the tendency for convective cur-
manikins (e.g. Fan and Keighley, 1991; Weder, et al., rents to form within air spaces as does whether the
1996). space is open to the ambient environment or en-
Air spaces 6–8 mm are associated with reduced closed. Differences in heat loss through layers is also
thermal resistance per unit thickness (demonstrated dependent on the orientation of the layers. For exam-
using a guarded hot plate apparatus) (Bomberg, ple, heat transfer in horizontal air layers has been
1992). Changes to the thermal resistance of the air shown to be greater than that occurring in vertical
space may be due to convection within the space (al- layers (Fowle, 1933; Spencer-Smith, 1977) and in in-
though losses have been considered negligible in air clined layers 700 (Hollands, et al., 1976). Fowle
spaces less than 10–13 mm thick) (Gretton, et al., (1933) provided some insight into the behavior of
1998; Spencer-Smith, 1977). Air spaces 13 mm narrow, vertical layers of air (presumably within en-
have been linked to a significant reduction in thermal closed systems) when one vertical surface is heated.
resistance of assemblies of various types (e.g. infant Heat transfer through horizontal air spaces was re-
bedding (Wilson, et al., 1999a; Wilson, et al., 2000), portedly twice that of vertical ones for air spaces up
single- and multiple-layer arrangements of thin fab- to 12 mm (Fowle, 1933). The rate of convective heat
rics (Cain and Farnworth, 1986) and unspecified sur- transfer was independent of the thickness of the layer
faces (Fowle, 1933)). (or distance between the surfaces) but proportional to
the square of the temperature difference between the
Thickness, orientation and height of air layers heated and unheated surfaces and also to the cube of
Most published test methods used to measure ther- the distance between the two surfaces (Fowle, 1933).
mal resistance of fabrics attempt to minimize the air At a critical flow speed (which occurred above 20°C
space (International Organization for Standardiza- temperature difference and a width of 12 mm) turbu-
tion, 1993(E)) or specify a standard air space (e.g. an lence affected the relationship. This is consistent with
air space of 13 mm (Gretton, et al., 1998)) in an at- the onset of convection observed by Cain and Farn-
tempt to control the variability when air spaces are worth (1986) at an air space thickness of approxi-
incorporated into the test structure. The effects of in- mately 13 mm. Fowle (1933) considered smooth ver-
troducing either standard or non-controlled air layers, tical surfaces, and a closed system was implied
both small (0.2–6.3 mm) (e.g. Belding, et al., 1947; (which may account for the differences between these
Gretton, et al., 1998; Kerslake, 1991) and large (up to findings and those from several studies on textiles in
atmosphere) (e.g. Fujimoto, et al., 1989; Kerslake, which heat transfer through horizontal air spaces has
1991; Ruckman, 1997; Wilson, et al., 1999a; Wilson, been more common).
et al., 2000; Woo, et al., 1994) have been investi- Spencer-Smith (1977) also reported greater heat
gated, presumably because of the wish to better ap- loss in horizontal than vertical air layers between fab-
proximate real conditions. Results suggest thermal rics, using a hot plate and a range of different fabric
resistance of the fabric and air space differs accord- types (i.e. linen drill, cotton mesh, worsted and
ing to the characteristics of the air gap. For example, blends of ‘Leavil’ and wool). Differences increased
gaps may be irregularly shaped, of varying thickness with increasing difference between the surface and
and/or distributed in varying ways throughout the as- the air temperatures (Differences range from 10–
sembly (Wilson, et al., 1999a; Wilson, et al., 2000), 45°C). Spencer-Smith (1977) appears to summarize
orientated horizontally or vertically (Fowle, 1933; published findings on the movement of heat through
Hollands, et al., 1976; Spencer-Smith, 1977), and/or clothing under dry steady-state conditions providing
open or closed to the ambient environment (Sat- a useful summary of the effect of air layers and their
sumoto, et al., 1991). orientation on thermal resistance. Equations for cal-
Wilson, et al., (1999a) suggested convection in culating heat loss through clothing layers (in the ab-
very large air space, over a critical thickness of hori- sence of moisture) under differing conditions are
zontal air space, may reduce thermal resistance given, although the extent to which they have been
(‘dry’) per unit thickness of the air space. A maxi- validated is unclear.
mum thermal resistance at approximately 20mm was Hollands et. al. (1976) investigated the effect of
reached and then declined slightly as the thickness of varying the Rayleigh number (sub-critical to 105) of
72 C. A. Wilson, R. M. Laing and D. J. Carr

horizontal and inclined air layers (0–700) with a large than when standing. Compression of clothing over
aspect ratio and which were heated from below. A the back, seat and thigh by the weight of the body or
theoretical result was compared to the experimental the clothing itself, may account for some of these ob-
data. The theoretical solution predicted Nusselt num- served differences (Havenith, et al., 1990). Nielsen,
bers in good agreement (maximum error approxi- et al., (1985) suggested the increase in thermal resist-
mately 5%) for all experimental data except that ob- ance resulted from the boundary layer being thicker
tained for air layers inclined at 700. (Note the theoret- in the region of much of the body (i.e. seated com-
ical result was confirmed experimentally for air lay- pared to standing) reflecting the Grashof number.
ers inclined at 150 only and the materials used to Whether the thermal resistance of the clothing is al-
form the confining surfaces were metallic rather than tered measurably by changes in the boundary layer is
fabric). Theory suggested that if the fluid flow was not entirely clear. The findings of Nielsen, et al.
independent of the x direction, the Nusselt number (1985) confirm earlier work by Belding, et al., (1947)
was a function of the relationship Ra cos f (where, and Olesen, et al., (1982).
Ra is the Rayleigh number and f the angle of the in-
clined air layer). Fit
The distribution and thickness of air layers and the
2. Multiple layer assemblies ratio of air to fiber depend on the fit and design of
Garments and garment design garments. Geometric factors are associated with fit
Differences in the thermal resistance of garments also affect thermal resistance because fit affects the
have been related to differences in fit (i.e. the size of extent, thickness, and distribution of air trapped.
the air space) (Fan and Keighley, 1991; Havenith, et Radiation differences as a result of garment fit have
al., 1990), and garment length and the amount of also been identified (Kakitsuba, et al., 1987). The
body surface covered by garments (McCullough, et volume of enclosed air also changes with differences
al., 1983). Wearing garments increases the external in the fit of the assembly. Differences in microclimate
surface area and therefore theoretically increases the volume within infant bedding have been identified.
capacity for heat loss. In addition, the rate of micro- The extent of differences in microclimate volume
climate air exchanged with the surrounding atmos- was affected by the product type and the fit/wrapping
phere using different designs of lightweight cloth arrangement (Holland, et al., 1999). Weder et al.,
coats has shown that although designs may differ (1996) showed (using both a sweating and moving
greatly, the conditions of use (especially the func- arm and a sweating guarded hotplate) that wide
tional position of the arm) may result in comparable sleeves provided on average 30% greater thermal re-
microclimate volumes and exchange rates (Shivers, et sistance than narrow, more fitted sleeves. A greater
al., 1977). Thus, surface configurations and fabric reduction in ‘dry’ and ‘wet’ thermal resistance during
arrangements affect the thermal resistance of the gar- movement in narrow compared to wide sleeves was
ments. Differences may be comparatively small also found (Weder, et al., 1996).
(Goldman, 1977; Spencer-Smith, 1977), but increase Literature on the effects of use/fit on the size,
in importance as the thickness of the assembly, and shape and distribution of air layers within three-di-
thus the surface area available for heat exchange, in- mensional products is sparse. Wilson, et al., (1999b)
creases. measured and documented the formation of air layers
in infant bedding during use and Havenith, et al.
Posture (1990) attempted to identify differences due to fit by
Changes in posture alter the contact between body comparing lean and obese persons wearing the same,
surfaces and thus affect the surface area available for identically designed and sized garments (the gar-
heat exchange and the relative importance of the vari- ments therefore tighter and looser for the obese and
ous modes of heat transmission (Wheldon, 1982). lean persons respectively). Tight clothing resulted in
The posture of a wearer/user also affects the arrange- 6–31% lower thermal resistance than clothing which
ment of the material layers and air spaces. When was loosely fitted (Havenith, et al., 1990). The differ-
standing (with garments hanging freely) there are dif- ences were reduced when an additional impermeable
ferences in the extent and pattern of air trapping. For rain cover of coated nylon was used, perhaps because
example, a slight increase in total thermal resistance the air space under the garment collapsed under the
of ‘light’ clothing and a slight decrease for ‘heavy’ weight of the added garment or because of a reduc-
clothing has been observed when sitting compared to tion in the air and water vapor permeability of the
standing (Lotens, 1989). Nielsen, et al., (1985) and material. The number of people included in the study
Havenith, et al., (1990) showed the intrinsic thermal was also small (n2 lean and 2 obese), so what these
resistance of clothing was 6–18% less when sitting findings mean is not entirely clear.
Air and Air Spaces—the Invisible Addition to Thermal Resistance 73

Lotens (1989) suggested that the effect of fit (pos- less air reportedly passes through open closures than
ture and other variables) were predictable (from occurred with wider spaces (19 mm: (3/4 inch)) (Fon-
mathematical calculations) and proposed that the in- seca and Breckenridge, 1965). Spacing that mini-
trinsic thermal resistance of the loosely-and tightly- mized the effects of air penetration through the fabric
fitted clothing did not necessarily differ. Rather, the were less effective when air movement was through
space formed below the outer garment contributed to an open closure. Differences in penetration through
differences in total thermal resistance (Lotens, 1989). closures when the underlying space was narrow com-
From a practical perspective though, Fan (1989) re- pared to large were attributed to increased resistance
ported that loose clothing seemed to be more affected to air flow associated with narrow air spaces and re-
by wind and body movement than tight clothing. duced penetration of air in wind. Yet where air move-
However, this study was conducted using a hollow ment occurs due to forced bellowing as a result of
cylinder and the representative clothing assembly movement lower thermal resistance was reported in
evaluated (two-layer fabric combinations) are not di- narrow than wide sleeves (Weder, et al., 1996). Thus
rectly comparable with those investigated by Lotens the source of air movement and the type of fabric
(1989). may affect thermal resistance derived from the under-
lying air spaces.
Closures and vents
The rate and volume of air exchanged between the 3. Effects of wind and body movement
assembly microclimate and the ambient environment Convective heat flow
is affected by the extent of air flow through the fabric Air movement in and around the clothed person
itself and/or through openings in the garment (Ruck- has been referred to as: i) bellows ventilation and
man, et al., 1999). Many examples of the effects of pumping (i.e. the circulation of air through the cloth-
vents have been published between 1980–2000 ap- ing as a result of rhythmic movement such as walk-
proximately. The effect of sleeve, side and sleeve and ing) (Ruckman, et al., 1999; Vokac, et al., 1973); ii)
side openings on the thermal resistance of two walk- a ‘chimney effect’ (Goldman, 1977; Vokac, et al.,
ing jackets (PTFE laminated and polyurethane 1973; Weder, et al., 1996); and/or iii) ambient air
coated) worn by six persons simulating fell-walking motion (Vokac, et al., 1973). Whether there is suffi-
in a controlled environmental chamber was investi- cient evidence for the existence of three separate
gated by Ruckman (1999). At rest, in a non-venti- ‘processes’ is unclear. Irrespective of the mode of air
lated condition, both jackets performed similarly. movement exchange of garment microclimate “air”
Under these conditions the type of fabric had a (generally warmed and partly or fully saturated with
greater effect on the thermal resistance than the use water vapor) with that of the ambient environment
of ventilation, although during exercise, ventilation occurs through garments. An investigation into dif-
was the primary factor. Thermal resistance was most ferences between black and white robes worn in hot
affected by the position of the pit zip as evidenced by climates demonstrated the effectiveness of heat trans-
the change in skin temperature. Pit zip openings at fer resulting from air movement through assemblies
both the sleeve and side seams had the greatest effect as a case study (n1)(Shkolnik, et al., 1980). In spite
followed by sleeve openings, side openings and no of additional heat absorbed by black robes (indicated
openings. Weder, et al., (1996) also found a strong by the surface temperature of 47°C compared to
fabric effect when examining ‘wet’ thermal resist- 41°C in white robes), total heat gain did not differ be-
ance of wide and narrow sleeved garments in the ab- tween black or white robes. Skin temperature under
sence of arm movement. both robes was similar 37°C. Enhanced convective
There is fairly clear evidence that air flow through transfer of heat in the air space under the black robe
vents depends partly on the extent to which the fabric either by enhanced bellows and/or chimney ventila-
is permeable to air (as demonstrated by Weder (1996) tion was the explanation provided.
examining sleeves, and Reischl and Stransky (1980)
and Bridgman et al., (1990) in garment assemblies Variables affecting heat loss during air movement
for firefighters). Similar results have been reported Variables affecting the rate of heat loss and the
with impermeable materials as well (e.g. Holland, et magnitude of air exchange due to air movement have
al., (2002) in industrial helmets with a variety of vent been attributed to many factors including: i) fabric
arrangements; and Weder, et al., (1996) in sleeves). properties, such as thickness, weight, fiber arrange-
The effect of openings may be countered by the re- ment, structure, permeability to air (Babus’Haq, et
sistance of underlying air spaces adjacent to or under al., 1996; Epps and Song, 1992; Shoshani, et al.,
the opening. Where smaller narrower air spaces (3 1993) and product type (Holland, et al., 1999), ii)
mm: (1/8 inch)) were formed over a heated cylinder, temperature differences between the skin and ambi-
74 C. A. Wilson, R. M. Laing and D. J. Carr

ent air (Goldman, 1977), iii) clothing design, such as to wind (Fan, 1998; Fan and Keighley, 1989) (the
the number of layers with low air permeability, fit and magnitude of the effect was described as ‘minimal’).
drape, fiber density, flexibility of layers and adequacy Air permeability of fabrics and convection was also
of closures (Babus’Haq, et al., 1996; Breckenridge, considered important in still air when vertical air
1977; Shivers, et al., 1977), and iv) mode and condi- spaces were 10 mm (Satsumoto, et al., 1991).
tions of use such as tucking method, posture, body Under these circumstances convection may occur in
motion, and wind (Crockford, et al., 1972; Havenith, the air space because the heated air in the air space is
et al., 1990; Holland, et al., 1999; Lotens and of lower air pressure than that of the ambient envi-
Havenith, 1988). ronment. The permeability of the materials to air is
Differences in thermal resistance among various therefore important as this property affects air move-
fiber types when exposed to wind has been attributed ment through the material and with a consequential
to the presence of loose fibers on the surface of the effect on thermal resistance. This has been shown ir-
fabric (e.g. cotton compared to viscose or polyester) respective of whether the lower perimeter of the air
which act to extend the thickness of the boundary air space was open or closed. In wider spaces (i.e. 19
layer (Babus’Haq, et al., 1996). Thus, the thermal re- mm), permeability of the material to air was impor-
sistance of fabrics with raised surfaces are likely to tant only if the air space was closed (Satsumoto, et
decrease more when exposed to wind than fabrics al., 1991), possibly due to reduced or negligible air
lacking fibrous extensions, resulting from erosion of flow through the wider air space than would occur if
the thicker air layer with increasing wind speed. the bottom of the space was open.
However, in the study by Babus’Haq, et al. (1996) the Changing some properties of underlying layers in-
cotton, viscose and polyester fabrics were selected as cluding the distribution of air layers also affects ther-
summer-appropriate textiles and seem not to have mal resistance as shown by Wilson, et al. (1999a;
been matched for other properties such as weight, 2000) using bedding materials. A single large space
density or thickness. Complete information on fabric was less thermally resistant than a number of smaller
structure was not provided. air spaces of comparable total thickness. This differ-
In knit fabrics, the major reduction in thermal re- ence may be due to the greater convective heat losses
sistance has been linked to wind speeds up to 2 m/s. and bellows when fewer larger air spaces are formed
Higher velocities had a minimal effect on the thermal than when there are a greater number of smaller air
resistance of knitted fabrics of various fiber types spaces. Increasing the thickness of inner layer fabrics
(acrylic and wool), and stitch densities (39.11 to 63.5 also resulted in significant changes to thermal resist-
loops/cm2) and structures (n17) (Shoshani and ance on the windward side of a hollow cylinder cov-
Shaltiel, 1993). Whether these findings apply to a ered in an inner batting fabric and an outer fabric
wider range of knit structures is unclear. However, re- (Fan, 1998; Fan and Keighley, 1989). Thermal resist-
sults are consistent with those of Fan and Keighley ance increased by 200% when the thickness of inner
(1989) who suggested that once the microclimate and layer fabrics was increased from 5 to 40 mm (Fan,
surface air was disturbed, further air movement may 1998; Fan and Keighley, 1989).
have only a reduced effect on thermal resistance.
Two layers of the same fabric were more thermally Body motion and wind
resistant than one layer. However, this effect was not The boundary layer, clothing, and hence the total
additive, probably reflecting the greater proportion of thermal resistance of an assembly are maximized in
total thermal resistance accounted for by the bound- still air without body movement (Fan and Keighley,
ary air layer in one-layer arrangements (Wilson, et 1991). In wind, the thermal resistance of the bound-
al., 1999a; Wilson, et al., 2000). The work of Wilson, ary air layer has been shown to decrease by 56% (de-
et al. (1999a; 2000) was carried out under standard termined using a thermal manikin) possibly due to
conditions (20°C, 65% R.H. and 35°C, 40% R.H for changes in the pattern of convective heat loss due to
‘dry’ and ‘wet’ thermal resistance respectively) and increased air velocity (Olesen, et al., 1982).
an air speed of 1 m/s. At air speeds greater than 6 However, the effect of body movement and wind
m/s, the thermal resistance of two (and also three) velocity on intrinsic clothing, boundary air layer and
layers approached that of one layer (Babus’Haq, et total thermal resistance suggests that body movement
al., 1996) suggesting that the effectiveness of layered and wind velocity cannot be considered in isolation.
textiles in wind may be affected by the air permeabil- For example, air movement has a minimal or no ef-
ity of the outer layer. fect on the intrinsic thermal resistance of clothing
Altering the air permeability of inner fabric layers when the wearer is standing in light wind (1.1m/s)
has been shown to affect the thermal resistance of (Breckenridge, 1977; Nielsen, et al., 1985). Instead
fabrics wrapped around a cylindrical model exposed of affecting the intrinsic insulation of the clothing,
Air and Air Spaces—the Invisible Addition to Thermal Resistance 75

wind strips away the boundary air layer (Forbes, ley, 1989; Fonseca and Breckenridge, 1965). Body
1968) and thus decreases total thermal resistance. movement and wind may also introduce forced air
Such a decrease reportedly occurs irrespective of flow and thus increase convection within air spaces.
whether it is the effect of body or air movement that Forced air flow varies, depending on the characteris-
is being examined (Goldman, 1988; Holcombe, tics of the underlying layers. When the air space is
1984). narrow, resistance to penetration of air on the wind-
Motion (cycling and walking) and its effect on the ward side is greatest and more air is forced into mid-
thermal resistance of clothing and the boundary air dle layers than when resistance to air movement is
layer were compared by Lotens (1989) and Nielsen, lower (Fonseca and Breckenridge, 1965).
et al., (1985). The intrinsic thermal resistance of the Wind penetration into air gaps in the garment re-
clothing and the boundary air layer were both higher sults in changes from laminar to turbulent flow as the
when cycling (at 40 rpm 20 W) than when walking size of the space or temperature gradient increases
(3.6 km/h) (Nielsen, et al., 1985) and both lower than (Spencer-Smith, 1977). The presence of such air
when standing. Tests were conducted in a controlled movement, possibly also associated with turbulent air
chamber where the chambers mean environmental flow due to body movement (Hassenboehler and
temperature was not specified but reported as equal Vigo, 1982; Kerslake, 1991), may decrease the ther-
to ambient temperature, relative humidity was ap- mal efficiency of an assembly.
proximately 50% and air movement 0.05 m/s. The
thermal resistance of clothing, boundary layer and Conclusions
total resistance was measured on human participants The size and shape (and thus volume), and the dis-
using indirect calorimetry. Differences were attrib- tribution of air layers affect thermal resistance of
uted to the different patterns of body movement typi- three-dimensional fabric arrangements such as gar-
cal of the two activities and possibly also to posture. ments, when they are being used. The posture and
Walking moves air layers external to the garment, body movement of the wearer, the relationship of gar-
disturbs air enclosed between the body of the wearer ment dimension to those of the body, and product de-
and the garment, and reduces the thickness of the sign, as well as wind velocity also affect thermal re-
boundary air layer. Walking in light wind (1.1 m/s) sistance by modifying the characteristics of those air
reduces the thermal resistance of both intrinsic ther- spaces. An apparently simple concept, this review
mal resistance of clothing (by 35%) and the thermal shows how complex the variables affecting thermal
resistance of the boundary air layer (by 63%) resistance are, and highlights the care required when
(Nielsen, et al., 1985) (estimated from dry heat loss, attempting to examine the contribution which each
mean skin, clothing and ambient air temperature). A makes to total thermal resistance.
similar decrease in intrinsic thermal resistance of
clothing when walking in still air has been observed, References
suggesting body movement has a larger affect on in- Babus’Haq, R. F., Hiasat, M. A. A. and Probert, S. D.
trinsic thermal resistance of clothing than increasing ‘Thermal Insulating behaviour of single and multiple
air velocity (Nielsen, et al., 1985). layers of textiles under wind assault’, Applied Energy, 54
Increasing wind speed, reduces thermal resistance (4), 375–391, 1996.
of both the clothing and the boundary layer (Fan and Belding, H. S., Russell, H. D., Darling, R. C. and Folk, G.
E. ‘Analysis of factors concerned in maintaining energy
Keighley, 1991). The thermal resistance of a nude
balance for dressed men in extreme cold: effects of activ-
and clothed manikin was determined and total, intrin-
ity on the protective value and comfort of an Arctic uni-
sic, and boundary thermal resistance calculated for form’, The American Journal of Physiology, 149, 223–
the clothing. The effect of wind on the thermal resist- 239, 1947.
ance of clothing seems to be overshadowed by the ef- Bo, Q. and Nakajima, T. ‘A numerical study on the heat
fect of body movement while wind has a greater ef- loss from clothed humans—effects of air space and
fect on thermal resistance of the boundary layer than clothing properties’, Journal of the Human-Environment
movement does (Havenith, et al., 1990). Consistent System, 5 (1), 33–40, 2001.
with the effect of walking, wind modifies the air Bogaty, H., Hollies, N. R. S., Hintermaier, J. C. and Harris,
spaces within and over the assembly surface. The M. ‘The nature of a fabric surface: thickness-pressure re-
lationships’, Textile Research Journal, 23 (2), 108–114,
magnitude of the wind effect appears to depend on
1953.
the air permeability of wind-breaking layers, the ef- Bomberg, M. (1992). ‘Application of three ASTM test
fectiveness of closures, and whether or not the fabric methods to measure thermal resistance of clothing’. In
layers were spaced so air could pass freely through Performance of protective clothing: ASTM STP 1133,
them, and escape, without penetrating deeply into the Fourth volume. Edited by McBriarty, J. P. and Henry, N.
clothing system (Breckenridge, 1977; Fan and Keigh- W. Philadelphia, American Society for Testing and Mate-
76 C. A. Wilson, R. M. Laing and D. J. Carr

rials, 399–426. Journal, 52 (8), 510–517, 1982.


Breckenridge, J. R. (1977). ‘Effects of body motion on con- Havenith, G. ‘Heat balance when wearing protective cloth-
vective and evaporative heat exchanges through various ing’, Annals of Occupational Hygiene, 43 (5), 289–296,
designs of clothing’. In Clothing comfort. Interaction of 1999.
thermal, ventilation, construction and assessment fac- Havenith, G., Heus, R. and Lotens, W. A. ‘Resultant cloth-
tors, Edited by Hollies, N. R. S. and Goldman, R. F. ing insulation: a function of body movement, posture,
Michigan, Ann Arbor Science Publishers Inc., 153–166. wind, clothing fit and ensemble thickness’, Ergonomics,
Bridgman, L. E. 1990. The effect of vents on garment mi- 33 (1), 67–84, 1990.
croclimate air exchange. Unpublished Honours Disserta- Holcombe, B. ‘The thermal insulation performance of tex-
tion, University of Otago, Dunedin, New Zealand. tile fabrics’, Wool Science Review, 60 (April), 12–22,
Cain, B. and Farnworth, B. ‘Two new techniques for deter- 1984.
mining the thermal radiative properties of thin fabrics’, Holcombe, B. V., Brooks, J. H., Schneider, A. M. and Watt,
Journal of Thermal Insulation, 9 (April), 301–322, 1986. I. C. (1988). ‘The objective measurement of clothing
Crockford, G. W., Crowder, M. and Prestidge, S. P. ‘A trace comfort’. In The Textile Institute 1988 Annual World
gas technique for measuring clothing microclimate air Conference, Sydney, The Textile Institute, 436–445.
exchange rates’, British Journal of Industrial Medicine, Holcombe, B. V. and Hoschke, B. N. ‘Dry heat transfer
29 (4), 378–386, 1972. characteristics of underwear fabrics’, Textile Research
Epps, H. H. and Song, M. K. ‘Thermal transmittance and Journal, 53 (6), 368–375, 1983.
air permeability of plain weave fabrics’, Clothing and Holland, E., J., Laing, R. M., Lemmon, T. L. and Niven, B.
Textiles Research Journal, 11 (1), 10–17, 1992. E. ‘Helmet design to facilitate thermoneutrality during
Fan, J. ‘Heat transfer through clothing assemblies in windy forestry harvesting’, Ergonomics, (In press) 2002.
conditions’, Textile Asia, 29 (10), 39–45, 1998. Holland, E. J., Wilson, C. A., Laing, R. M. and Niven, B. E.
Fan, J. and Keighley, J. H. ‘Theoretical and experimental ‘Microclimate ventilation of infant bedding’, Interna-
study of thermal insulation of clothing in windy condi- tional Journal of Clothing Science and Technology, 11
tions’, International Journal of Clothing Science and (4), 226–239, 1999.
Technology, 1 (1), 21–29, 1989. Hollands, K. G. T., Raithby, G. D. and Konicek, L. ‘Corre-
Fan, J. and Keighley, J. H. ‘An investigation on the effects lation equations for free convective heat transfer in hori-
of: body motion, clothing design and environmental con- zontal layers of air and water’, International Journal of
ditions by using a fabric manikin’, International Journal Heat and Mass Transfer, 18 (7–8), 879–884, 1975.
of Clothing Science and Technology, 3 (5), 6–13, 1991. Hollands, K. G. T., Unny, T. E., Raithby, G. D. and Kon-
Fonseca, G. F. and Breckenridge, J. R. ‘Wind penetration icek, L. ‘Free convective heat transfer across inclined air
through fabric systems. Part I’, Textile Research Journal, layers’, Journal of Heat Transfer, 98 (2), 189–193, 1976.
35 (2), 95–103, 1965. International Organization for Standardization, 1993(E).
Forbes, W. H. (1968). ‘Laboratory and field studies. Gen- ISO 11092: Textiles—Physiological effects—Measure-
eral principles’. In Physiology of heat regulation and the ment of thermal and water-vapour resistance under
science of clothing, Edited by Newburgh, L. H. New steady-state conditions (sweating guarded-hotplate test).
York, United States of America, Hefner Publishing Co., Geneve: International Organization for Standardization.
320–325. Kakitsuba, N., Michna, H. and Mekjavic, I. B. ‘Clothing
Fowle, F. E. (Ed.). (1933). Smithsonian physical tables. surface area as related to body volume and clothing mi-
Washington, Smithsonian Institute, 327. croenvironment volume’, Aviation, Space, and Environ-
Fujimoto, T., Niwa, M. and Seki, N. ‘Effect of an air layer mental Medicine, 58 (May), 411–416, 1987.
on thermal insulation of clothing of materials’, Journal Kerslake, D. M. ‘The insulation provided by infants bed-
of the Textile Machinery Society of Japan, 35 (4), 28, clothes’, Ergonomics, 34 (7), 893–907, 1991.
1989. Laing, R. M. and Sleivert, G. G. Clothing, textiles, and
Goldman, R. F. (1977). ‘Thermal comfort factors: concepts human performance. Textile Progress, 32 (2), 1–122,
and definitions. Invited review paper’. In Clothing com- 2002.
fort. Interaction of thermal, ventilation, construction and Lotens, W. A. ‘The actual insulation of multilayer cloth-
assessment factors, Edited by Hollies, N. R. S. and Gold- ing’, Scandinavian Journal of Work Environment &
man, R. F. Michigan, Ann Arbor Science Publishers Inc., Health, 15 (Suppl. 1), 66–75, 1989.
3–8. Lotens, W. A. and Havenith, G. (1988). ‘Ventilation of rain-
Goldman, R. F. (1988). ‘Standards for human exposure to wear determined by a trace gas method’. In Environmen-
heat’. In Environmental ergonomics, Edited by Mek- tal ergonomics: sustaining human performance in harsh
javic, I. B., Banister, E. W. and Morrison, J. B. London, environments, Edited by Mekjavic, I. B., Banister, E. W.
Taylor and Francis Ltd., 99–136. and Morrison, J. B. London, Taylor and Francis Ltd.
Gretton, J. C., Brook, D. B., Dyson, H. M. and Harlock, S. McCullough, E. A., Jones, B. W. and Zbikowski, P. J. ‘The
C. ‘Moisture vapor transport through waterproof breath- effect of garment design on the thermal insulation values
able fabrics and clothing systems under a temperature of clothing’, American Society of Heating, Refrigerating
gradient’, Textile Research Journal, 68 (12), 936–941, and Air-Conditioning Engineers, 89 (2A), 327–352,
1998. 1983.
Hassenboehler, C. B. and Vigo, T. L. ‘A mixed flow ther- Morris, G. J. ‘Thermal properties of textile materials’,
mal transmittance tester for textiles’, Textile Research Journal of the Textile Institute Transactions, 44 (10),
Air and Air Spaces—the Invisible Addition to Thermal Resistance 77

T449–T477, 1953. ing comfort. Interaction of thermal, ventilation, con-


Morris, M. A. ‘Thermal insulation of single and multiple struction and assessment factors, Edited by Hollies, N.
layers of fabrics’, Textile Research Journal, 25 (9), R. S. and Goldman, R. F. Michigan, Ann Arbor Science
766–773, 1955. Publishers, 167–181.
Nielsen, R., Olesen, B. W. and Fanger, P. O. ‘Effect of phys- Shkolnik, A., Taylor, C. R., Finch, V. and Borut, A. ‘Why
ical activity and air velocity on the thermal insulation of do Bedouins wear black robes in hot deserts?’, Nature,
clothing’, Ergonomics, 28 (12), 1617–1631, 1985. 283, 373–375, 1980.
Obendorf, S. K. and Smith, J. P. ‘Heat transfer characteris- Shoshani, Y. and Shaltiel, S. ‘Effect of wind velocity on the
tics of nonwoven insulating materials’, Textile Research thermal resistance of knitted fabrics’, Knitting Tech-
Journal, 56 (11), 691–696, 1986. nique, 15 (2), 91–93, 1993.
Olesen, B. W., Sliwinska, E., Madsen, T. L. and Fanger, P. Spencer-Smith, J. L. ‘The physical basis of clothing com-
O. ‘Effect of body posture and activity on the thermal in- fort. Part 2: heat transfer through dry clothing assem-
sulation of clothing: measurements by a movable thermal blies’, Clothing Research Journal, 5 (1), 3–17, 1977.
manikin’, ASHRAE Transactions, 88 (4), 791–805, Ukponmwan, J. O. The thermal-insulation properties of
1982. fabrics. Textile Progress, 24 (4), 1–57, 1993.
Parker, S. P. (Ed.). (1994). McGraw-Hill dictionary of sci- Vokac, Z., Køpke, V. and Keül, P. ‘Assessment and analysis
entific and technical terms (5th ed.). New York, Mc- of the bellows ventilation of clothing’, Textile Research
Graw-Hill Inc. Journal, 43 (8), 474–482, 1973.
Peirce, F. T. and Rees, W. H. ‘The transmission of heat Weder, M. S., Zimmerli, T. and Rossi, R. M. (1996). ‘A
through textile fabrics—Part II’, Journal of The Textile sweating and moving arm for the measurement of ther-
Institute, 37 (9), T181–T204, 1946. mal insulation and water vapour resistance of clothing’.
Reischl, U. and Stransky, A. ‘Assessment of ventilation In Performance of protective clothing. ASTM STP 1237,
characteristics of standard and prototype fire fighter pro- Fifth volume. Edited by Johnson, J. S. and Mansdorf, S.
tective clothing’, Textile Research Journal, 50 (3), Z. West Conshohocken, American Society for Testing
193–201, 1980. and Materials, 257–268.
Ruckman, J. E. ‘Water vapour transfer in waterproof Wheldon, A. E. ‘Energy balance in the newborn baby: use
breathable fabrics. Part 1: under steady state conditions’, of a manikin to estimate radiant and convective heat
International Journal of Clothing Science and Technol- loss’, Physics in Medicine and Biology, 27 (2), 285–296,
ogy, 9 (1), 10–22, 1997. 1982.
Ruckman, J. E., Murray, R. and Choi, H. S. ‘Engineering of Wilson, C. A., Laing, R. M. and Niven, B. E. ‘Estimating
clothing systems for improved thermophysiological com- thermal resistance of multiple-layer bedding materials -
fort. The effect of openings’, International Journal of re-examining the problem’, Journal of the Human-Envi-
Clothing Science and Technology, 11 (1), 37–52, 1999. ronment System, 2 (1), 69–85, 1999a.
Satsumoto, Y., Takeuchi, M. and Ishikawa, K. ‘The effect Wilson, C. A., Laing, R. M. and Niven, B. E. ‘Multiple-
of air permeability and radiation property on free layer bedding materials and the effect of air spaces on
convective heat transfer of clothing system’, Sen-I ‘wet’ thermal resistance of dry materials’, Journal of the
Gakkaishi, 47 (6), 263–270, 1991. Human-Environment System, 4 (1), 23–32, 2000.
Schneider, A. M. and Holcombe, B. V. (1988). ‘The role of Wilson, C. A., Niven, B. E. and Laing, R. M. ‘Estimating
radiation in fabric warmth’. In Advanced workshop on thermal resistance of bedding from thickness of materi-
the application of mathematics and physics in the wool als’, International Journal of Clothing Science and Tech-
industry proceeding, Edited by Carnaby, G. A., Wood, E. nology, 11 (5), 262–276, 1999b.
J. and Story, L. F. Lincoln, New Zealand, 488–502. Woo, S. S., Shalev, I. and Barker, R. L. ‘Heat and moisture
Shivers, J. L., Yeh, K., Fourt, L. and Spivak, S. M. (1977). transfer through nonwoven fabrics. Part II: moisture dif-
‘The effects of design and degree of closure on microcli- fusivity’, Textile Research Journal, 64 (4), 190–197,
mate air exchange in lightweight cloth coats’. In Cloth- 1994.

También podría gustarte