Computational Modeling of Protective Clothing

Published on December 2016 | Categories: Documents | Downloads: 35 | Comments: 0 | Views: 200
of 10
Download PDF   Embed   Report

Computational Modeling ofProtective Clothing

Comments

Content

ORIGINAL PAPER/PEER-REVIEWED

Computational Modeling of
Protective Clothing
By James J. Barry, Principal Engineer, and Roger W. Hill, Engineer, Creare Inc.

ABSTRACT
Models based on computational fluid dynamics (CFD) have
been developed to predict the performance of chemical and
steam/fire protective clothing. The software computes the
diffusive and convective transport of heat and gases/vapors;
capillary transport of liquids; vapor and liquid sorption phenomena and phase change; and the variable properties of the
various clothing layers. It can also model the effects of sweating and humidity transport to help assess the thermal stress
imposed on the wearer of the clothing. Specialized geometry/grid representations of clothed humans have been created for performing two- and three-dimensional simulations.
Comparisons with experimental data show good agreement
in predicting the effects of fiber swell due to transients in
humidity, and the models have been used to predict the sensitivity of clothing performance to material properties such as
permeability under varying environmental conditions.
Applications of the models include analysis of chemical protective garment design for military and emergency response
personnel, comparisons of thermally protective materials for
steam or fire protection, and evaluation of clothing test data.
NOMENCLATURE
cp
constant pressure specific heat (J/kg-K)
h
enthalpy (J/kg)
∆hv enthalpy of vaporization (J/kg)
J
species diffusion flux (kg/m2-s)
keff
effective thermal conductivity (W/m-K)
K
intrinsic permeability (m2)
m
mass fraction

mm
mass source per unit volume (kg/m3-sec)
Q
enthalpy of desorption from solid phase (J/kg)
p
pressure (N/m2)
25 INJ Fall 2003

Pc
R
S
s
t
T
v

capillary pressure (N/m2)
fiber regain (kgbw/kgds)
source term
saturation
time (s)
temperature (K)
velocity (m/s)

GREEK SYMBOLS
ε
volume fraction (m3 of quantity/m3)
µ
dynamic viscosity (N-s/m3)
ρ
density (kg of quantity/m3 of quantity)
SUBSCRIPTS
β
liquid-phase
γ
gas phase
σ
solid phase
bl
bound liquid
ds
dry solid
lv
liquid-to-vapor
ls
liquid-to-solid
sat
saturation
sv
solid-to-vapor
v
vapor
INTRODUCTION
Protective clothing provides laboratory and hazardous
materials workers, fire fighters, military personnel, and others
with the means to control their exposure to chemicals, biological materials, and heat sources. Depending on the specific
application, the textile materials used in protective clothing
must provide high performance in a number of areas, including impermeability to hazardous chemicals, breathability,

light weight, low cost, and ruggedness. Nonwoven
textiles provide key components of an increasing number of these protective garments.
Development of protective materials presently relies
heavily on testing. Swatches of textile materials undergo laboratory tests to measure their properties, and the
performance of partial or complete clothing products
are measured in test chambers or field tests. The objective of the work reported here is development of computational models for predicting the performance of
textile materials in protective applications. Such models complement testing by enabling property data from
tests of textile swatches to be used in predictions of
integrated multilayer garments under varying environmental conditions.
Computational fluid dynamics (CFD) provides the
basis for the models. In CFD, software solves the govFigure 1
erning equations for mass, momentum, and heat trans- TRANSPORT AND PHASE CHANGE
fer in fluids over a two- or three-dimensional computaFABRIC MODEL
tional mesh. The calculations result in predictions of
the flow velocity, temperature, pressure, and composition at each location in the mesh. Though commercial CFD
software provides many built-in capabilities, it does not offer
all of the physical models required to address the complex
Vapor continuity equation:
multiphase processes that occur in textile fabrics. As
described in this paper, detailed models for these processes
have been developed, integrated into widely used commercial
software, and used for a series of validation and applications
calculations.
FABRIC MODEL FORMULATION
The fabric model simulates the transport of a liquid- and
vapor-phase fluid that can undergo phase change (e.g., water)
and an inert gas (air) in a textile layer, as illustrated in Figure
1. Several new models and capabilities were added to a standard commercial CFD code (FLUENT Version 6.0, Fluent Inc.,
Lebanon, NH). These capabilities include:
• Vapor phase transport (variable permeability)
• Liquid phase transport (wicking)
• Fabric property dependence on moisture content
• Vapor/liquid phase change (evaporation/condensation)
• Sorption to fabric fibers
In the fabric, transport equations are derived for mass,
momentum, and energy in the gas and liquid phases by volume-averaging techniques (Gibson; 1994, 1996). Definitions
for intrinsic phase average, global phase average, and spatial
average for porous media are those given by Whitaker (1977,
1998). Since the fabric porosity is not constant due to changing amounts of liquid and bound water, the source term for
each transport equation includes quantities that arise due to
the variable porosity. These equations are summarized in general form below.
Gas phase continuity equation:

PROCESSES IN

(1b)

(2a)

(2b)

(2c)

Gas phase momentum equation:
(3a)

(3b)
Liquid transport:
(4a)

(4b)
(1a)

26 INJ Fall 2003

The model equations are recast into a form compatible with
the base CFD code, then implemented using the software’s “user-defined function” (UDF) capability. An
extension to the CFD software’s graphical user interface
(5a)
(GUI) provides the means to adjust fabric parameters.
Figure 2 illustrates the integration of the models with the
CFD software.

Energy equation (combined):

MULTIDIMENSIONAL REPRESENTATIONS OF
CLOTHED HUMANS
To simulate clothing performance, computational representations of clothed humans have been developed using
geometry and grid generation tools. To elucidate the basic
physics, a simple 2-D cylinder—sized to mimic a human
arm—is clothed with one or more layers of fabric, Figure 3.
The fabric layers may vary in number, thickness, and concentricity. The relatively small size of the mesh (about 14,000
cells) enables calculations to be completed quickly, allowing
for significant numbers of sensitivity runs.
Figures 4 and 5 depict more complex 3-D models of an arm
and torso, respectively. Two layers of fabric clothe the arm.
The undulations visible on the clothing surface near the inner
elbow are in the outer layer of fabric only. The torso model
here is clad in a single fabric layer, a crew-neck T-shirt. Both

(5b)

(5c)

In addition to the transport equations, a substantial body of
supporting equations for properties and interphase exchange
rates are included. Permeability is estimated using experimentally determined permeabilities at dry and saturated conditions assuming that the flow resistance is proportional to the
regain:

(6)

Relative permeability constitutive relationships are based
on saturation (Wang and Beckerman, 1993; Perre et. al., 1993).
Effective thermal conductivity is computed by the method of
Progelhof et al. (1976). Capillary pressure is represented by
the Leverett J-function form (Udell, 1984; Wang and
Beckerman, 1993).

Figure 2
INTEGRATION OF FABRIC MODELS WITH
CFD SOFTWARE

27 INJ Fall 2003

Figure 3
2-D MODEL OF A SIMPLIFIED ARM: (A)
WITH UNIFORM GAP AND TWO CLOTHING
LAYERS; (B) GEOMETRY WITH SINGLE
CLOTHING LAYER AND NONUNIFORM GAP

Figure 4
3-D MODEL OF ARM WITH TWO CLOTHING
LAYERS

Figure 5
3-D MODEL OF TORSO (CLAD IN T-SHIRT)

arm and torso models are based on laser scans of humans.
Scanned points are brought into computer-aided design software for creation of the body surface and generation of clothing layers. The geometry is then exported to the CFD software’s preprocessor for grid generation. The arm and torso
models have approximately 0.2 million and 1.2 million grid
points, respectively.
Figure 6 illustrates a 3-D representation of a soldier. This
representation models only flow over the outer surface of the
clothing and protective gear rather than the heat and mass
transfer within the fabrics themselves. Its primary purpose is
predicting the flow field immediately around the soldier
(including recirculation zones downstream) for assessing
impingement of wind-driven chemical agents. This computational representation was developed from a commercial
Viewpoint Premier 3-D digital image that was imported into

28 INJ Fall 2003

Figure 6
3-D MODEL OF KNEELING SOLDIER IN
PROTECTIVE GEAR
the CFD grid generation software for simplification (i.e.,
removal of excessive detail), creation of an enclosed volume,
and meshing. The resulting computational mesh uses
approximately 1.6 million grid points.
COMPUTATIONAL RESULTS
Using the gridded human representations and the CFD
software with integrated fabric physics models, a wide range
of calculations can be performed. The sections below give
several illustrative examples, including some validation comparisons with experimental data.
Steady-State Data Comparison for Dynamic Moisture
Permeation Cell. A single piece of a cotton fabric sample

Relative humidity at the bottom outlet

Figure 7
RELATIVE HUMIDITY AT THE DMPC
LOWER CHANNEL EXIT (CENTER) VERSUS
PRESSURE DIFFERENCE BETWEEN THE
UPPER AND LOWER CHANNELS
(0.384 mm thick) is placed in the center of an experimental
apparatus known as the DMPC (Dynamic Moisture
Permeation Cell) described in Gibson (1996) and Gibson et al.
(1995). In this cell, a small rectangular swatch of fabric is
clamped in place, and controlled flows of gas and/or vapor
flow parallel to its upper and lower surfaces. For the data
considered here, inlet nitrogen flows at 0.57 m/s above and
below the sample. The upper flow is at 100% relative humidity, while the lower flow is at 0% relative humidity. Pressures
in the upper and lower flow streams are set to impose a pressure difference across the sample, so that gas is forced through
the fabric.
Figure 7 provides a plot of the relative humidity at the bottom outlet as a function of the pressure difference across the
fabric sample. Positive ∆P indicates a condition where a higher percentage of flow leaves the lower channel exit.
Numerical results are provided for three cases: the fabric layer
thickness discretized with one, two, or eight cells. All sets of
the computed results follow the experimental data trend with
pressure drop closely. The minor deviations from the experimental data are attributed to possible small differences
between the modeling input parameters and the experimental
conditions, actual fabric flow resistance that is not linearly
proportional to the fabric regain, and fabric regain that deviates from the assumed regain function with relative humidity.
The ability to obtain accurate results even for very coarse discretizations of the fabric is important, since fine discretizations can make large 3-D simulations impractically expensive
to perform.
Transient Data Comparisons for Dynamic Moisture
Permeation Cell. Also using a model of a DMPC, comparisons were made of the predicted temperature response in the
center of the fabric samples with transient DMPC experimental thermocouple measurements. In these tests, the DMPC
was loaded with two layers of a given fabric (cotton, nylon,

29 INJ Fall 2003

Table 1
SEQUENCE OF CHANGES IN RELATIVE
HUMIDITY FOR DMPC TRANSIENTS
State
Initial Condition
Transient 1
Transient 2
Transient 3
Transient 4

Inlet Flow
Relative Humidity
0% above and below
100% above and below
0% above and below
60% above, 0% below
80% above, 0% below

silk, polyester, or wool) with a thermocouple located between
the layers to record temperature. In the cases considered here,
inlet flow rates of 0.57 m/s of nitrogen at 20ºC are again supplied above and below the fabric samples. The humidity of
the inlet flows is varied with time to provide transient conditions.
Starting at t = 0 sec, the system undergoes a series of four
step changes in the relative humidity of the inlet flows as
shown in Table 1. The switches in relative humidity occur at
30-minute intervals (15 minutes for polyester). Increases in
the relative humidity induce sorption of water from the moist
N2 into the dry fibers and a release of the latent heat and heat
of sorption as the fibers approach equilibrium with relative
humidity conditions of the incoming gas. Decreases in relative humidity induce desorption of the bound water which
requires thermal energy to supply the latent heat and heat of
sorption. (Note that for these transients no free liquid is present.) Consequently, step increases in relative humidity produce an initial rise in fabric temperature while step decreases
produce an initial fall in fabric temperature. An effective diffusion coefficient for absorption of moisture into the fabric
fiber was determined by fitting to data for the first peak in
humidity (Transient 1) and used without change for the full
transient calculation.
Figure 8 compares predicted temperatures with the DMPC
experimental data for cotton, nylon, silk, polyester, and wool.
The magnitude and trends of the simulated temperatures
agree well with the experimental data for all of the materials
and inlet condition step changes. The largest deviations
between model and experiment are seen for large humidity
changes for wool, where the relaxation of the temperature
back after the sorption-induced spike is slower in the simulations.
Flow Over 2-D Arm Clothed in Nonwoven Material.
Simulations were performed using the 2-D model of a simplified arm shown in Figure 3a using a single protective clothing
layer of 130 µm thick Tyvek®. The layer was placed such that
a 1.1 cm uniform gap was present between the arm and the
Tyvek inner surface. The Tyvek was assumed to have a permeability of 1.52x10-14 m2 based on experimental air permeability measurements. From water vapor permeability data,
the diffusion coefficient for vapor transport through the Tyvek
layer was estimated to be 300 times smaller than for diffusion

Figure 8
COMPARISON OF SIMULATED AND EXPERIMENTAL TEMPERATURES
FOR VARIOUS CLOTH SAMPLES IN THE DMPC
of water vapor in air. Simulations were performed as transients, with initial conditions of a temperature of 300 K and
humidity of zero. At the beginning of the transient, a 2.5 mph
dry wind was imposed on the system. To mimic conditions of
a resting metabolic load, a sweat flux of 1.5x10-5 kg/m2 and
heat flux of 60 W/m2 was applied at the surface of the arm.
For this simulation, thermal radiation was included since it is
the dominant mode of heat transfer in the gap between the
arm and the clothing material due to the low permeability of
Tyvek.
Figure 9 shows the contours of temperature at 5, 25, and 75
seconds into the transient. Contours of the water vapor mass
fraction are shown in Figure 10 for the same times. The surface of the arm is observed to quickly reach a temperature of
about 303 K. The evaporation of sweat increases the water
vapor mass fraction in the gap layer at a constant rate until the
air becomes saturated at the arm surface at a time of about 22
30 INJ Fall 2003

seconds. After this time, the sweat flux and heat flux remain
the same, but a decreasing amount of the sweat evaporates
(the balance accumulates as liquid on the skin surface) causing the temperature at the arm surface to rise until it reaches
a temperature of about 308 K. (Note that the water vapor
mass fraction plots use a log scale to enhance visualization of
the concentration outside the fabric layer. The flow structure
observed to the right of the cylinders is due to the vortex
shedding.) A fabric with a lower resistance to air flow
through the fabric or vapor diffusion through the fabric
would result in a lower equilibrium temperature.
3-D Model of Human Torso. Figure 11 illustrates a simple
example of heat transfer from a 3-D model of a human torso.
Using the computational mesh of Figure 5, a 5 mph wind at
300K and 70% relative humidity is imposed on the front of the
torso, with skin conditions set to reflect a moderate workload:
a sweat flux of 3x10-5 kg/m2-s and heat flux of 100 W/m2. The

Figure 9
TEMPERATURE CONTOURS FOR FLOW
OVER A SIMPLIFIED 2-D ARM WITH A SINGLE
PROTECTIVE LAYER

Figure 10
WATER VAPOR MASS FRACTION
CONTOURS FOR FLOW OVER A SIMPLIFIED 2D ARM WITH A SINGLE PROTECTIVE LAYER

T-shirt is modeled as 1 mm thick cotton fabric. Figure 11a
shows the temperature at the skin surface and velocity vectors
in a plane around the torso for conditions of full closure
between the layer of clothing and skin at the bottom of the
shirt, ends of the sleeves, and at the neck (i.e., snug fit at neck,
sleeves, and waist). Figure 11b and 11c show the temperature
at a slice slightly below the armpit for the same conditions
and for conditions wherein the closures are open (i.e., loose fit
at neck, sleeves, and waist).
As shown in the figures, the highest temperatures are under
the arm. Furthermore, the presence of the closures and their

effect of limiting the ability of flow from the environment to
enter the area under the shirt results in higher temperatures
(as expected).
Protection From Chemical Agents. The simplified twodimensional model of an arm (Figure 3) has been used to
explore in detail the variation of agent exposure at the arm to
variables such as wind speed, location of the liquid droplet on
the clothing surface, and gaps between clothing and skin. The
evaporation and transport of vapor from a liquid agent
droplet on a clothing surface is modeled as a small region of
the clothing with sufficient liquid agent present such that the

31 INJ Fall 2003

Figure 12
SCHEMATIC OF BASE GEOMETRY FOR
TWO-DIMENSIONAL SIMULATIONS OF
EVAPORATING SURFACE AGENT (WITH
DROPLET LOCATIONS SHOWN)

Figure 11
SIMULATION RESULTS OF CLOTHED 3-D
HUMAN TORSO: (A) TEMPERATURES AT SKIN
SURFACE AND FLOW AROUND TORSO WITH
COMPLETE CLOSURE (SNUG FIT) AT NECK,
SLEEVES, AND WAIST; (B) TEMPERATURE
NEAR ARM PIT FOR COMPLETE CLOSURE; (C)
TEMPERATURE NEAR ARM PIT FOR OPEN
CONDITIONS (LOOSE FIT) AT NECK, SLEEVES,
AND WAIST.

air in this region remains saturated with agent vapor (as
determined by the vapor pressure of the liquid agent).
Transport of the liquid agent itself is neglected (i.e., it cannot
wick and does not occupy volume) and the system is assumed
isothermal. This simplified approach models agent vapor carried away by the flow over and through the clothing layer but
does not track the rate of evaporation or time to dissipate a
droplet of known size.
Figure 12 shows the base geometry: a 10 cm diameter arm
covered with a single 0.5 mm thick clothing layer having a 1.1
cm gap between the surface of the arm and the clothing layer.
Fabric permeability was assumed to be a cotton shell (permeability approximately 2.4x10-12 m2) without chemical absorbent
materials. Vapor properties correspond to GB (Sarin). The
transient simulations were performed with a time step chosen
to resolve the shedding frequency of the familiar von Karman
vortex street for flow over a circular cylinder. Two wind
speeds (5 and 20 mph) were simulated.
The following sets of geometric parameters were considered:
• Base Geometry with an incident wind direction of 0o, 45o,
90o, 135o, and 180o relative to droplet. Condition of 0o corresponds to droplet located at stagnation point.
• Wind direction at 0o relative to droplet with single clothing layer having uniform gap spacings of 0.2, 0.6, 1.1, and 2.1
cm.
• Wind direction at 0o relative to droplet with single clothing layer eccentric to the arm with non-uniform gap spacing
of 4.1:1, and 6.8:1 (maximum/minimum of 1.767 cm/0.433 cm

Agent Concentration (mg/cm3)

relatively small effect on agent exposure
for droplets located at the stagnation point.
Reduction of the gap near the droplet,
however, can significantly increase exposure.
Figure 13 illustrates these results for the
baseline geometry. The average concentration of agent at the skin surface is highest
when the agent droplet is located at the
stagnation point, and higher wind speeds
enhance this effect. When droplets are
located elsewhere, however, the increase in
dilution caused by the higher wind speeds
reduces the average concentration.
Liquid Wicking in a Nonwoven Wipe. A
thermally bonded composite nonwoven
material of 70% polypropylene and 30%
cellulose was used in numerical simulations of a simple wicking experiment. In
the experiment, the material was suspended vertically and lowered until the bottom
Figure 13
edge was submerged in a pool of liquid
AVERAGE AGENT CONCENTRATIONS AT THE SKIN
water at approximately 20°C. The water
(BASELINE GEOMETRY) AS FUNCTIONS OF WIND SPEED
immediately began to wick upward into the
AND AGENT DROPLET LOCATIONS
sample, and the process was recorded on
videotape for subsequent extraction of the
wicking
height
as
a function of time. Figure 13 shows the
and 1.918 cm/0.282 cm, respectively with minimum thickness
located at the stagnation point). The 4.1:1 geometry is shown experimentally observed liquid height in the sample above
the free surface of the liquid as a function of time. The experin Figure 3b.
• Wind direction at 0o relative to droplet with two clothing iment was repeated four times with different samples with no
layers (Figure 3a) having uniform gap spacings. First case significant difference observed in the wicking behavior.
To perform the computational simulations of these tests,
with inner surface of inner and outer layers located 0.55 cm
and 1.1 cm from arm surface, respectively. Second case with several key material properties that characterize the nonwoinner surface of inner and outer layers located 1.1 cm and 2.1 ven sample were needed. Sample thickness measurements
along with air flow and pressure drop measurements through
cm from arm surface, respectively.
All simulations were run until the air flow reached a sta- a sample were used to determine the intrinsic permeability.
tionary oscillatory state. Agent concentration was assessed to Absorption of water into the bound state was neglected, a
determine the maximum concentration and the area-averaged very reasonable assumption for the polypropylene fibers but
concentration over the arm surface. The results of the simula- an area for refinement in treatment of the cellulose fibers,
tions indicate that the predicted agent exposure depends on a especially for more accurate treatment of long-time behavior.
competition between penetration of agent through the cloth- The porosity was estimated (not measured) to be 70% for the
ing (which causes higher exposure) and dilution of agent calculations. The gas and liquid relative permeabilities were
vapor (higher air penetration reduces exposure concentra- assumed to be third order functions of the relative saturation.
The computational domain included both the vertical nonwotions). Specific observations include:
• Agent droplets located at the stagnation point result in the ven sample and an adjacent gas region over which a very low
velocity saturated and isothermal air flow was forced (resulthighest exposures.
• At 5 mph wind speed, diffusion remains a major transport ing in negligible evaporation effects). At t = 0s, a liquid satumechanism under the clothing. Agent exposures are the same ration, s, of 0.99 was imposed at one end of the composite
order of magnitude for all orientations. The stagnation point sample in which an irreducible saturation of 0.1 was assumed,
orientation, i.e., 0o, gives 2 to 3 times higher exposures than and the transport equations (excluding the energy equation)
were integrated in time. (The liquid saturation is defined as
the other orientations.
• For 20 mph wind speeds, convection becomes more dom- the ratio of the liquid volume fraction to the sum of the liquid
inant, and dilution becomes a significant effect. Resulting and gas volume fractions.)
Figure 14 shows the liquid height as a function of time for
exposures are one to two orders of magnitude lower than for
5 mph for all orientations except 0o (stagnation point). At the two cases (the height was defined as the location where the
liquid saturation was approximately equal to the irreducible
stagnation point, exposures increase at high wind speeds.
• Increasing gap width and multiple clothing layers have a saturation). Given uncertainties in the nonwoven properties,
33 INJ Fall 2003

ACKNOWLEDGEMENTS
This work was supported by the U.S. Army
Soldier Biological Chemical Command
under contract DAAD16-00-C-9255.

Wicking Height

References
P.W. Gibson, “Governing Equations for
Multiphase Heat and Mass Transfer in
Hygroscopic
Porous
Media
with
Applications to Clothing Materials,”
Technical Report Natick/TR-95/004 (U.S.
Army Natick Research, Development, and
Engineering Center, Natick, MA, 1994).
Gibson, P. W., “Multiphase Heat and Mass
Transfer Through Hygroscopic Porous
Media with Applications to Clothing
Materials,” Technical Report Natick/TR97/005, U.S. Army Natick Research,
Development and Engineering Center,
Natick, MA, 1996.
Figure 14
Gibson, P. W., Kendrick, C., Rivin, D.,
Sicuranza,
L., Charmchi, M., “An
LIQUID HEIGHT VS. TIME FOR WATER WICKING
Automated
Water
Vapor Diffusion Test
VERTICALLY INTO A NONWOVEN COMPOSITE
Method for Fabrics, Laminates, and Films,”
the numerical simulations show very good agreement with Journal of Coated Fabrics, Vol. 24, 1995, pp. 322-345.
Perre, P., Moser, M., and Martin, M., “Advances in
the experimental observations. The simulation assuming a
capillary pressure function based on 100% of the water sur- Transport Phenomena During Convective Drying with
face tension value overpredicts the wicking height. However, Superheated Steam and Moist Air,” International Journal of
decreasing the capillary pressure function by a factor of two Heat and Mass Transfer, Vol. 36, 1993, pp. 2725-2746.
Progelhof, R., Throne, J., Ruetsch, R., “Methods for
brings the numerical solution into much closer agreement.
The reduction is capillary pressure function could be inter- Predicting the Thermal Conductivity of Composite Systems:
preted as the presence of a non-zero contact angle in the com- A Review,” Polymer Engineering and Science, Vol. 16, 1976, pp.
posite, a plausible situation considering that the material is 615-625.
Udell, K. S., “Heat Transfer in Porous Media Considering
70% polypropylene and initially dry.
Phase Change and Capillarity—The Heat Pipe Effect,”
International Journal of Heat and Mass Transfer, Vol. 28, 1984, pp.
CONCLUSION
In the development of protective clothing and other textiles, 485-495.
Wang, C. Y. and Beckerman, C., “A Two-Phase Mixture
modeling offers a powerful companion to experiments and
testing. Detailed models for vapor and liquid phase transport Model of Liquid-Gas Flow and Heat Transfer in Capillary
within textile fabrics have been developed and integrated Porous Media—I. Formulation,” International Journal of Heat
with CFD software. Validation of the models with experi- and Mass Transfer, Vol. 36, pp. 2747-2758, 1993.
S. Whitaker, in Advances in Heat Transfer, Vol. 13, edited by
mental data has been successful for moisture absorption, permeability, and wicking. Validation with additional data for J. Hartnett, (Academic Press, New York, 1977) p. 119.
S. Whitaker, in Advances in Heat Transfer, Vol. 31, edited by
more varied conditions is needed. Applications of the soft— INJ
ware thus far have included analysis of chemical penetration J. Hartnett, (Academic Press, New York, 1998) p. 1.
of garments due to wind, investigation of flow in swatch-testing equipment, and effects of seal leakage on protective clothing performance. Future applications could involve assessment of thermal comfort/stress on wearers of protective
clothing, effects of layering on protective performance, and
sensitivity to textile permeability and wicking properties.
New opportunities for validation and application of the modeling tools are sought.

Return to Table of Contents
34 INJ Fall 2003

Sponsor Documents

Or use your account on DocShare.tips

Hide

Forgot your password?

Or register your new account on DocShare.tips

Hide

Lost your password? Please enter your email address. You will receive a link to create a new password.

Back to log-in

Close