10
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 phe- nomena and phase change; and the variable properties of the various clothing layers. It can also model the effects of sweat- ing and humidity transport to help assess the thermal stress imposed on the wearer of the clothing. Specialized geome- try/grid representations of clothed humans have been creat- ed 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 sen- sitivity of clothing performance to material properties such as permeability under varying environmental conditions. Applications of the models include analysis of chemical pro- tective 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 c p constant pressure specific heat (J/kg-K) h enthalpy (J/kg) h v enthalpy of vaporization (J/kg) J species diffusion flux (kg/m 2 -s) k eff effective thermal conductivity (W/m-K) K intrinsic permeability (m 2 ) m mass fraction m m mass source per unit volume (kg/m 3 -sec) Q enthalpy of desorption from solid phase (J/kg) p pressure (N/m 2 ) P c capillary pressure (N/m 2 ) R fiber regain (kg bw /kg ds ) S source term s saturation t time (s) T temperature (K) v velocity (m/s) GREEK SYMBOLS ε volume fraction (m 3 of quantity/m 3 ) µ dynamic viscosity (N-s/m 3 ) ρ density (kg of quantity/m 3 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, biolog- ical 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, includ- ing impermeability to hazardous chemicals, breathability, Computational Modeling of Protective Clothing By James J. Barry, Principal Engineer, and Roger W. Hill, Engineer, Creare Inc. ORIGINAL PAPER/PEER-REVIEWED 25 INJ Fall 2003

Computational Modeling of Protective Clothing

Embed Size (px)

DESCRIPTION

Computational Modeling ofProtective Clothing

Citation preview

Page 1: Computational Modeling of Protective Clothing

ABSTRACTModels based on computational fluid dynamics (CFD) have

been developed to predict the performance of chemical andsteam/fire protective clothing. The software computes thediffusive and convective transport of heat and gases/vapors;capillary transport of liquids; vapor and liquid sorption phe-nomena and phase change; and the variable properties of thevarious clothing layers. It can also model the effects of sweat-ing and humidity transport to help assess the thermal stressimposed on the wearer of the clothing. Specialized geome-try/grid representations of clothed humans have been creat-ed for performing two- and three-dimensional simulations.Comparisons with experimental data show good agreementin predicting the effects of fiber swell due to transients inhumidity, and the models have been used to predict the sen-sitivity of clothing performance to material properties such aspermeability under varying environmental conditions.Applications of the models include analysis of chemical pro-tective garment design for military and emergency responsepersonnel, comparisons of thermally protective materials forsteam or fire protection, and evaluation of clothing test data.

NOMENCLATUREcp 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 fractionmm mass source per unit volume (kg/m3-sec)Q enthalpy of desorption from solid phase (J/kg)p pressure (N/m2)

Pc capillary pressure (N/m2)R fiber regain (kgbw/kgds)S source terms saturationt time (s)T temperature (K)v velocity (m/s)

GREEK SYMBOLSε volume fraction (m3 of quantity/m3)µ dynamic viscosity (N-s/m3)ρ density (kg of quantity/m3 of quantity)

SUBSCRIPTSβ liquid-phaseγ gas phaseσ solid phasebl bound liquidds dry solidlv liquid-to-vaporls liquid-to-solidsat saturationsv solid-to-vaporv vapor

INTRODUCTIONProtective clothing provides laboratory and hazardous

materials workers, fire fighters, military personnel, and otherswith the means to control their exposure to chemicals, biolog-ical materials, and heat sources. Depending on the specificapplication, the textile materials used in protective clothingmust provide high performance in a number of areas, includ-ing impermeability to hazardous chemicals, breathability,

Computational Modeling ofProtective ClothingBy James J. Barry, Principal Engineer, and Roger W. Hill, Engineer, Creare Inc.

ORIGINAL PAPER/PEER-REVIEWED

25 INJ Fall 2003

Page 2: Computational Modeling of Protective Clothing

light weight, low cost, and ruggedness. Nonwoventextiles provide key components of an increasing num-ber of these protective garments.

Development of protective materials presently reliesheavily on testing. Swatches of textile materials under-go laboratory tests to measure their properties, and theperformance of partial or complete clothing productsare measured in test chambers or field tests. The objec-tive of the work reported here is development of com-putational models for predicting the performance oftextile materials in protective applications. Such mod-els complement testing by enabling property data fromtests of textile swatches to be used in predictions ofintegrated multilayer garments under varying envi-ronmental conditions.

Computational fluid dynamics (CFD) provides thebasis for the models. In CFD, software solves the gov-erning equations for mass, momentum, and heat trans-fer in fluids over a two- or three-dimensional computa-tional mesh. The calculations result in predictions ofthe flow velocity, temperature, pressure, and composi-tion at each location in the mesh. Though commercial CFDsoftware provides many built-in capabilities, it does not offerall of the physical models required to address the complexmultiphase processes that occur in textile fabrics. Asdescribed in this paper, detailed models for these processeshave been developed, integrated into widely used commercialsoftware, and used for a series of validation and applicationscalculations.

FABRIC MODEL FORMULATIONThe 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 Figure1. Several new models and capabilities were added to a stan-dard 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 fibersIn the fabric, transport equations are derived for mass,

momentum, and energy in the gas and liquid phases by vol-ume-averaging techniques (Gibson; 1994, 1996). Definitionsfor intrinsic phase average, global phase average, and spatialaverage for porous media are those given by Whitaker (1977,1998). Since the fabric porosity is not constant due to chang-ing amounts of liquid and bound water, the source term foreach transport equation includes quantities that arise due tothe variable porosity. These equations are summarized in gen-eral form below.

Gas phase continuity equation:

Vapor continuity equation:

Gas phase momentum equation:

Liquid transport:

26 INJ Fall 2003

Figure 1TRANSPORT AND PHASE CHANGE PROCESSES IN

FABRIC MODEL

(1a)

(1b)

(2b)

(2c)

(3a)

(3b)

(4a)

(4b)

(2a)

Page 3: Computational Modeling of Protective Clothing

Energy equation (combined):

In addition to the transport equations, a substantial body ofsupporting equations for properties and interphase exchangerates are included. Permeability is estimated using experi-mentally determined permeabilities at dry and saturated con-ditions assuming that the flow resistance is proportional to theregain:

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

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

MULTIDIMENSIONAL REPRESENTATIONS OFCLOTHED HUMANSTo simulate clothing performance, computational repre-

sentations of clothed humans have been developed usinggeometry and grid generation tools. To elucidate the basicphysics, a simple 2-D cylinder—sized to mimic a humanarm—is clothed with one or more layers of fabric, Figure 3.The fabric layers may vary in number, thickness, and concen-tricity. The relatively small size of the mesh (about 14,000cells) enables calculations to be completed quickly, allowingfor significant numbers of sensitivity runs.

Figures 4 and 5 depict more complex 3-D models of an armand torso, respectively. Two layers of fabric clothe the arm.The undulations visible on the clothing surface near the innerelbow are in the outer layer of fabric only. The torso modelhere is clad in a single fabric layer, a crew-neck T-shirt. Both

27 INJ Fall 2003

(5a)

(5b)

(5c)

(6)

Figure 2INTEGRATION OF FABRIC MODELS WITH

CFD SOFTWARE

Figure 32-D MODEL OF A SIMPLIFIED ARM: (A)

WITH UNIFORM GAP AND TWO CLOTHINGLAYERS; (B) GEOMETRY WITH SINGLE

CLOTHING LAYER AND NONUNIFORM GAP

Page 4: Computational Modeling of Protective Clothing

arm and torso models are based on laser scans of humans.Scanned points are brought into computer-aided design soft-ware for creation of the body surface and generation of cloth-ing layers. The geometry is then exported to the CFD soft-ware’s preprocessor for grid generation. The arm and torsomodels have approximately 0.2 million and 1.2 million gridpoints, respectively.

Figure 6 illustrates a 3-D representation of a soldier. Thisrepresentation models only flow over the outer surface of theclothing and protective gear rather than the heat and masstransfer within the fabrics themselves. Its primary purpose ispredicting the flow field immediately around the soldier(including recirculation zones downstream) for assessingimpingement of wind-driven chemical agents. This computa-tional representation was developed from a commercialViewpoint Premier 3-D digital image that was imported into

the CFD grid generation software for simplification (i.e.,removal of excessive detail), creation of an enclosed volume,and meshing. The resulting computational mesh usesapproximately 1.6 million grid points.

COMPUTATIONAL RESULTSUsing the gridded human representations and the CFD

software with integrated fabric physics models, a wide rangeof calculations can be performed. The sections below giveseveral illustrative examples, including some validation com-parisons with experimental data.

Steady-State Data Comparison for Dynamic MoisturePermeation Cell. A single piece of a cotton fabric sample

28 INJ Fall 2003

Figure 43-D MODEL OF ARM WITH TWO CLOTHING

LAYERS

Figure 63-D MODEL OF KNEELING SOLDIER IN

PROTECTIVE GEAR

Figure 53-D MODEL OF TORSO (CLAD IN T-SHIRT)

Page 5: Computational Modeling of Protective Clothing

(0.384 mm thick) is placed in the center of an experimentalapparatus known as the DMPC (Dynamic MoisturePermeation Cell) described in Gibson (1996) and Gibson et al.(1995). In this cell, a small rectangular swatch of fabric isclamped in place, and controlled flows of gas and/or vaporflow parallel to its upper and lower surfaces. For the dataconsidered here, inlet nitrogen flows at 0.57 m/s above andbelow the sample. The upper flow is at 100% relative humid-ity, while the lower flow is at 0% relative humidity. Pressuresin the upper and lower flow streams are set to impose a pres-sure difference across the sample, so that gas is forced throughthe fabric.

Figure 7 provides a plot of the relative humidity at the bot-tom outlet as a function of the pressure difference across thefabric sample. Positive ∆P indicates a condition where a high-er percentage of flow leaves the lower channel exit.Numerical results are provided for three cases: the fabric layerthickness discretized with one, two, or eight cells. All sets ofthe computed results follow the experimental data trend withpressure drop closely. The minor deviations from the experi-mental data are attributed to possible small differencesbetween the modeling input parameters and the experimentalconditions, actual fabric flow resistance that is not linearlyproportional to the fabric regain, and fabric regain that devi-ates from the assumed regain function with relative humidity.The ability to obtain accurate results even for very coarse dis-cretizations of the fabric is important, since fine discretiza-tions can make large 3-D simulations impractically expensiveto perform.

Transient Data Comparisons for Dynamic MoisturePermeation Cell. Also using a model of a DMPC, compar-isons were made of the predicted temperature response in thecenter of the fabric samples with transient DMPC experimen-tal thermocouple measurements. In these tests, the DMPCwas loaded with two layers of a given fabric (cotton, nylon,

silk, polyester, or wool) with a thermocouple located betweenthe layers to record temperature. In the cases considered here,inlet flow rates of 0.57 m/s of nitrogen at 20ºC are again sup-plied above and below the fabric samples. The humidity ofthe inlet flows is varied with time to provide transient condi-tions.

Starting at t = 0 sec, the system undergoes a series of fourstep changes in the relative humidity of the inlet flows asshown in Table 1. The switches in relative humidity occur at30-minute intervals (15 minutes for polyester). Increases inthe relative humidity induce sorption of water from the moistN2 into the dry fibers and a release of the latent heat and heatof sorption as the fibers approach equilibrium with relativehumidity conditions of the incoming gas. Decreases in rela-tive humidity induce desorption of the bound water whichrequires thermal energy to supply the latent heat and heat ofsorption. (Note that for these transients no free liquid is pre-sent.) Consequently, step increases in relative humidity pro-duce an initial rise in fabric temperature while step decreasesproduce an initial fall in fabric temperature. An effective dif-fusion coefficient for absorption of moisture into the fabricfiber was determined by fitting to data for the first peak inhumidity (Transient 1) and used without change for the fulltransient calculation.

Figure 8 compares predicted temperatures with the DMPCexperimental data for cotton, nylon, silk, polyester, and wool.The magnitude and trends of the simulated temperaturesagree well with the experimental data for all of the materialsand inlet condition step changes. The largest deviationsbetween model and experiment are seen for large humiditychanges for wool, where the relaxation of the temperatureback after the sorption-induced spike is slower in the simula-tions.

Flow Over 2-D Arm Clothed in Nonwoven Material.Simulations were performed using the 2-D model of a simpli-fied arm shown in Figure 3a using a single protective clothinglayer of 130 µm thick Tyvek®. The layer was placed such thata 1.1 cm uniform gap was present between the arm and theTyvek inner surface. The Tyvek was assumed to have a per-meability of 1.52x10-14 m2 based on experimental air perme-ability measurements. From water vapor permeability data,the diffusion coefficient for vapor transport through the Tyveklayer was estimated to be 300 times smaller than for diffusion

29 INJ Fall 2003

Figure 7RELATIVE HUMIDITY AT THE DMPC

LOWER CHANNEL EXIT (CENTER) VERSUSPRESSURE DIFFERENCE BETWEEN THE

UPPER AND LOWER CHANNELS

Rel

ativ

e hu

mid

ity a

t the

bot

tom

out

let

Table 1SEQUENCE OF CHANGES IN RELATIVE

HUMIDITY FOR DMPC TRANSIENTS

Inlet FlowState Relative HumidityInitial Condition 0% above and belowTransient 1 100% above and belowTransient 2 0% above and belowTransient 3 60% above, 0% belowTransient 4 80% above, 0% below

Page 6: Computational Modeling of Protective Clothing

of water vapor in air. Simulations were performed as tran-sients, with initial conditions of a temperature of 300 K andhumidity of zero. At the beginning of the transient, a 2.5 mphdry wind was imposed on the system. To mimic conditions ofa resting metabolic load, a sweat flux of 1.5x10-5 kg/m2 andheat flux of 60 W/m2 was applied at the surface of the arm.For this simulation, thermal radiation was included since it isthe dominant mode of heat transfer in the gap between thearm and the clothing material due to the low permeability ofTyvek.

Figure 9 shows the contours of temperature at 5, 25, and 75seconds into the transient. Contours of the water vapor massfraction are shown in Figure 10 for the same times. The sur-face of the arm is observed to quickly reach a temperature ofabout 303 K. The evaporation of sweat increases the watervapor mass fraction in the gap layer at a constant rate until theair becomes saturated at the arm surface at a time of about 22

seconds. After this time, the sweat flux and heat flux remainthe same, but a decreasing amount of the sweat evaporates(the balance accumulates as liquid on the skin surface) caus-ing the temperature at the arm surface to rise until it reachesa temperature of about 308 K. (Note that the water vapormass fraction plots use a log scale to enhance visualization ofthe concentration outside the fabric layer. The flow structureobserved to the right of the cylinders is due to the vortexshedding.) A fabric with a lower resistance to air flowthrough the fabric or vapor diffusion through the fabricwould result in a lower equilibrium temperature.

3-D Model of Human Torso. Figure 11 illustrates a simpleexample of heat transfer from a 3-D model of a human torso.Using the computational mesh of Figure 5, a 5 mph wind at300K and 70% relative humidity is imposed on the front of thetorso, 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

30 INJ Fall 2003

Figure 8COMPARISON OF SIMULATED AND EXPERIMENTAL TEMPERATURES

FOR VARIOUS CLOTH SAMPLES IN THE DMPC

Page 7: Computational Modeling of Protective Clothing

T-shirt is modeled as 1 mm thick cotton fabric. Figure 11ashows the temperature at the skin surface and velocity vectorsin a plane around the torso for conditions of full closurebetween the layer of clothing and skin at the bottom of theshirt, ends of the sleeves, and at the neck (i.e., snug fit at neck,sleeves, and waist). Figure 11b and 11c show the temperatureat a slice slightly below the armpit for the same conditionsand for conditions wherein the closures are open (i.e., loose fitat neck, sleeves, and waist).

As shown in the figures, the highest temperatures are underthe arm. Furthermore, the presence of the closures and their

effect of limiting the ability of flow from the environment toenter the area under the shirt results in higher temperatures(as expected).

Protection From Chemical Agents. The simplified two-dimensional model of an arm (Figure 3) has been used toexplore in detail the variation of agent exposure at the arm tovariables such as wind speed, location of the liquid droplet onthe clothing surface, and gaps between clothing and skin. Theevaporation and transport of vapor from a liquid agentdroplet on a clothing surface is modeled as a small region ofthe clothing with sufficient liquid agent present such that the

31 INJ Fall 2003

Figure 9TEMPERATURE CONTOURS FOR FLOW

OVER A SIMPLIFIED 2-D ARM WITH A SINGLEPROTECTIVE LAYER

Figure 10WATER VAPOR MASS FRACTION

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

Page 8: Computational Modeling of Protective Clothing

air in this region remains saturated with agent vapor (asdetermined by the vapor pressure of the liquid agent).Transport of the liquid agent itself is neglected (i.e., it cannotwick and does not occupy volume) and the system is assumedisothermal. This simplified approach models agent vapor car-ried away by the flow over and through the clothing layer butdoes not track the rate of evaporation or time to dissipate adroplet of known size.

Figure 12 shows the base geometry: a 10 cm diameter armcovered with a single 0.5 mm thick clothing layer having a 1.1cm gap between the surface of the arm and the clothing layer.Fabric permeability was assumed to be a cotton shell (perme-ability approximately 2.4x10-12 m2) without chemical absorbentmaterials. Vapor properties correspond to GB (Sarin). Thetransient simulations were performed with a time step chosento resolve the shedding frequency of the familiar von Karmanvortex street for flow over a circular cylinder. Two windspeeds (5 and 20 mph) were simulated.

The following sets of geometric parameters were consid-ered:

• Base Geometry with an incident wind direction of 0o, 45o,90o, 135o, and 180o relative to droplet. Condition of 0o corre-sponds to droplet located at stagnation point.

• Wind direction at 0o relative to droplet with single cloth-ing layer having uniform gap spacings of 0.2, 0.6, 1.1, and 2.1cm.

• Wind direction at 0o relative to droplet with single cloth-ing layer eccentric to the arm with non-uniform gap spacingof 4.1:1, and 6.8:1 (maximum/minimum of 1.767 cm/0.433 cm

Figure 12SCHEMATIC OF BASE GEOMETRY FOR

TWO-DIMENSIONAL SIMULATIONS OFEVAPORATING SURFACE AGENT (WITH

DROPLET LOCATIONS SHOWN)

Figure 11SIMULATION RESULTS OF CLOTHED 3-D

HUMAN TORSO: (A) TEMPERATURES AT SKINSURFACE AND FLOW AROUND TORSO WITHCOMPLETE CLOSURE (SNUG FIT) AT NECK,

SLEEVES, AND WAIST; (B) TEMPERATURENEAR ARM PIT FOR COMPLETE CLOSURE; (C)

TEMPERATURE NEAR ARM PIT FOR OPENCONDITIONS (LOOSE FIT) AT NECK, SLEEVES,

AND WAIST.

Page 9: Computational Modeling of Protective Clothing

and 1.918 cm/0.282 cm, respectively with minimum thicknesslocated at the stagnation point). The 4.1:1 geometry is shownin Figure 3b.

• Wind direction at 0o relative to droplet with two clothinglayers (Figure 3a) having uniform gap spacings. First casewith inner surface of inner and outer layers located 0.55 cmand 1.1 cm from arm surface, respectively. Second case withinner surface of inner and outer layers located 1.1 cm and 2.1cm from arm surface, respectively.

All simulations were run until the air flow reached a sta-tionary oscillatory state. Agent concentration was assessed todetermine the maximum concentration and the area-averagedconcentration over the arm surface. The results of the simula-tions indicate that the predicted agent exposure depends on acompetition between penetration of agent through the cloth-ing (which causes higher exposure) and dilution of agentvapor (higher air penetration reduces exposure concentra-tions). Specific observations include:

• Agent droplets located at the stagnation point result in thehighest exposures.

• At 5 mph wind speed, diffusion remains a major transportmechanism under the clothing. Agent exposures are the sameorder of magnitude for all orientations. The stagnation pointorientation, i.e., 0o, gives 2 to 3 times higher exposures thanthe other orientations.

• For 20 mph wind speeds, convection becomes more dom-inant, and dilution becomes a significant effect. Resultingexposures are one to two orders of magnitude lower than for5 mph for all orientations except 0o (stagnation point). At thestagnation point, exposures increase at high wind speeds.

• Increasing gap width and multiple clothing layers have a

relatively small effect on agent exposurefor droplets located at the stagnation point.Reduction of the gap near the droplet,however, can significantly increase expo-sure.

Figure 13 illustrates these results for thebaseline geometry. The average concentra-tion of agent at the skin surface is highestwhen the agent droplet is located at thestagnation point, and higher wind speedsenhance this effect. When droplets arelocated elsewhere, however, the increase indilution caused by the higher wind speedsreduces the average concentration.

Liquid Wicking in a Nonwoven Wipe. Athermally bonded composite nonwovenmaterial of 70% polypropylene and 30%cellulose was used in numerical simula-tions of a simple wicking experiment. Inthe experiment, the material was suspend-ed vertically and lowered until the bottomedge was submerged in a pool of liquidwater at approximately 20°C. The waterimmediately began to wick upward into thesample, and the process was recorded onvideotape for subsequent extraction of the

wicking height as a function of time. Figure 13 shows theexperimentally observed liquid height in the sample abovethe free surface of the liquid as a function of time. The exper-iment was repeated four times with different samples with nosignificant difference observed in the wicking behavior.

To perform the computational simulations of these tests,several key material properties that characterize the nonwo-ven sample were needed. Sample thickness measurementsalong with air flow and pressure drop measurements througha sample were used to determine the intrinsic permeability.Absorption of water into the bound state was neglected, avery reasonable assumption for the polypropylene fibers butan area for refinement in treatment of the cellulose fibers,especially for more accurate treatment of long-time behavior.The porosity was estimated (not measured) to be 70% for thecalculations. The gas and liquid relative permeabilities wereassumed to be third order functions of the relative saturation.The computational domain included both the vertical nonwo-ven sample and an adjacent gas region over which a very lowvelocity saturated and isothermal air flow was forced (result-ing in negligible evaporation effects). At t = 0s, a liquid satu-ration, s, of 0.99 was imposed at one end of the compositesample in which an irreducible saturation of 0.1 was assumed,and the transport equations (excluding the energy equation)were integrated in time. (The liquid saturation is defined asthe ratio of the liquid volume fraction to the sum of the liquidand gas volume fractions.)

Figure 14 shows the liquid height as a function of time fortwo cases (the height was defined as the location where theliquid saturation was approximately equal to the irreduciblesaturation). Given uncertainties in the nonwoven properties,

33 INJ Fall 2003

Figure 13AVERAGE AGENT CONCENTRATIONS AT THE SKIN

(BASELINE GEOMETRY) AS FUNCTIONS OF WIND SPEEDAND AGENT DROPLET LOCATIONS

Age

nt C

once

ntra

tion

(mg/

cm3 )

Page 10: Computational Modeling of Protective Clothing

the numerical simulations show very good agreement withthe experimental observations. The simulation assuming acapillary pressure function based on 100% of the water sur-face tension value overpredicts the wicking height. However,decreasing the capillary pressure function by a factor of twobrings the numerical solution into much closer agreement.The reduction is capillary pressure function could be inter-preted as the presence of a non-zero contact angle in the com-posite, a plausible situation considering that the material is70% polypropylene and initially dry.

CONCLUSIONIn the development of protective clothing and other textiles,

modeling offers a powerful companion to experiments andtesting. Detailed models for vapor and liquid phase transportwithin textile fabrics have been developed and integratedwith CFD software. Validation of the models with experi-mental data has been successful for moisture absorption, per-meability, and wicking. Validation with additional data formore varied conditions is needed. Applications of the soft-ware thus far have included analysis of chemical penetrationof garments due to wind, investigation of flow in swatch-test-ing equipment, and effects of seal leakage on protective cloth-ing performance. Future applications could involve assess-ment of thermal comfort/stress on wearers of protectiveclothing, effects of layering on protective performance, andsensitivity to textile permeability and wicking properties.New opportunities for validation and application of the mod-eling tools are sought.

ACKNOWLEDGEMENTSThis work was supported by the U.S. ArmySoldier Biological Chemical Commandunder contract DAAD16-00-C-9255.

ReferencesP.W. Gibson, “Governing Equations for

Multiphase Heat and Mass Transfer inHygroscopic Porous Media withApplications to Clothing Materials,”Technical Report Natick/TR-95/004 (U.S.Army Natick Research, Development, andEngineering Center, Natick, MA, 1994).

Gibson, P. W., “Multiphase Heat and MassTransfer Through Hygroscopic PorousMedia with Applications to ClothingMaterials,” Technical Report Natick/TR-97/005, U.S. Army Natick Research,Development and Engineering Center,Natick, MA, 1996.

Gibson, P. W., Kendrick, C., Rivin, D.,Sicuranza, L., Charmchi, M., “AnAutomated Water Vapor Diffusion TestMethod for Fabrics, Laminates, and Films,”

Journal of Coated Fabrics, Vol. 24, 1995, pp. 322-345.Perre, P., Moser, M., and Martin, M., “Advances in

Transport Phenomena During Convective Drying withSuperheated Steam and Moist Air,” International Journal ofHeat and Mass Transfer, Vol. 36, 1993, pp. 2725-2746.

Progelhof, R., Throne, J., Ruetsch, R., “Methods forPredicting the Thermal Conductivity of Composite Systems:A Review,” Polymer Engineering and Science, Vol. 16, 1976, pp.615-625.

Udell, K. S., “Heat Transfer in Porous Media ConsideringPhase Change and Capillarity—The Heat Pipe Effect,”International Journal of Heat and Mass Transfer, Vol. 28, 1984, pp.485-495.

Wang, C. Y. and Beckerman, C., “A Two-Phase MixtureModel of Liquid-Gas Flow and Heat Transfer in CapillaryPorous Media—I. Formulation,” International Journal of Heatand Mass Transfer, Vol. 36, pp. 2747-2758, 1993.

S. Whitaker, in Advances in Heat Transfer, Vol. 13, edited byJ. Hartnett, (Academic Press, New York, 1977) p. 119.

S. Whitaker, in Advances in Heat Transfer, Vol. 31, edited byJ. Hartnett, (Academic Press, New York, 1998) p. 1. — INJ

34 INJ Fall 2003

Figure 14LIQUID HEIGHT VS. TIME FOR WATER WICKINGVERTICALLY INTO A NONWOVEN COMPOSITE

Wic

king

Hei

ght