adsorption modelling for building contaminant dispersal analysis

26
Indoor Air. 2.147-171 (1991) ~ ~~~~~~ 0 1991 Danish Technical Press, DK-Copenhagen Adsorption Modelling for Building Contaminant Dispersal Analysis James W. Axley* Massachusetts Institute of Technology, USA Abstract Two families of macroscopic adsorption models are for- mulated, based on f u n d a m t a l principles of ads- tion science and technology, that may be used for mac- roscopic (e.g., whole-building)contaminant dispersal analysis. The first family of adForprion mo&ls - the EquilibriumAdsorption (EA) Models - are based upon the simple requirement of adsorptiun equilibnum between adsorbent and room air. The second family - the Boundary Layer Diffusion Controlled Adsorption (BLDC) Models - add to the e q u i l h u m requirement a boundary layer model for d i f i i o n of the adsorbate from the room air to the adsorbent sutjiie. Two mem- bers of each of thesefamilies are explicitly discussed, one based on the linear adsqtion isotherm mo&l and the other on the Lungmuir isotherm model. The linear variants of each family are applied to model the ad- sorptiun dynamics of f m l d e h y d e in gypsum wall board and compared to measured data. These applica- tions and a more general cotlsideratiun of the dynamic character of adsorption provided by these modeh indi- cate that simple physical adsorption and &sorption transpoTtprocesses h e the potential to signifiantly affect the dispersal of contaminants in buildings. KEY WORDS: Indoor air quality, Contaminant dispersal analy- sis, Adsorption, Multi-zone modelling. Manuscript received: 16 November 1990 Accepted for publication: 25 January 1991 * James W. Axley, Building Technology Program, Massachusetts Institute of Technology, Cambridge, MA, USA Introduction The quality of air in buildings is clearly de- pendent on both the nature of air movement into, out of, and within building systems and the character and location of contaminant sources. As a result, researchers have focused on the complementary problems of the air- flow aspects of contaminant dispersal and the emission characteristics of contaminant sour- ces. It is also well recognized that some con- taminants, such as Radon-222 gas and its radioactive decay daughters, and nitrogen di- oxide undergo chemical or radiochemical de- composition, and modelling the dispersal of these contaminants demands the additional consideration of the chemical or radiochemi- cal mass transformation processes involved. It is also known that adsorption and de- sorption processes can affect the dispersal of contaminants in buildings. The evidence is pervasive. The lingering odours of winter clothes stored during the summer with naphthalene crystals, and the odour of clothes exposed to tobacco smoke or gasoline fumes that persists for days until washed are but two examples. Few would doubt that ad- sorption and desorption phenomena can and do play a role in the dispersal of airborne contaminants in buildings but, from the point of view of indoor air quality analysis, should we suspect these phenomena to be significant? Can adsorption and desorption significantly alter the dynamics of the disper- sal of indoor air contaminants? If so, under what circumstances will the effect be signifi- cant and how may one model the dispersal of

Upload: james-w-axley

Post on 21-Jul-2016

213 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Adsorption Modelling for Building Contaminant Dispersal Analysis

Indoor Air. 2.147-171 (1991) ~ ~~~~~~

0 1991 Danish Technical Press, DK-Copenhagen

Adsorption Modelling for Building Contaminant Dispersal Analysis James W. Axley* Massachusetts Institute of Technology, USA

Abstract Two families of macroscopic adsorption models are for- mulated, based on f undamta l principles of ads- tion science and technology, that may be used for mac- roscopic (e.g., whole-building) contaminant dispersal analysis. The first family of adForprion mo&ls - the Equilibrium Adsorption (EA) Models - are based upon the simple requirement of adsorptiun equilibnum between adsorbent and room air. The second family - the Boundary Layer Diffusion Controlled Adsorption (BLDC) Models - add to the equilhum requirement a boundary layer model for d i f i i on of the adsorbate from the room air to the adsorbent sutjiie. Two mem- bers of each of these families are explicitly discussed, one based on the linear adsq t ion isotherm mo&l and the other on the Lungmuir isotherm model. The linear variants of each family are applied to model the ad- sorptiun dynamics of f m l d e h y d e in gypsum wall board and compared to measured data. These applica- tions and a more general cotlsideratiun of the dynamic character of adsorption provided by these modeh indi- cate that simple physical adsorption and &sorption transpoTt processes h e the potential to signifiantly affect the dispersal of contaminants in buildings.

KEY WORDS: Indoor air quality, Contaminant dispersal analy- sis, Adsorption, Multi-zone modelling.

Manuscript received: 16 November 1990 Accepted for publication: 25 January 1991

* James W. Axley, Building Technology Program, Massachusetts Institute of Technology, Cambridge, MA, USA

Introduction The quality of air in buildings is clearly de- pendent on both the nature of air movement into, out of, and within building systems and the character and location of contaminant sources. As a result, researchers have focused on the complementary problems of the air- flow aspects of contaminant dispersal and the emission characteristics of contaminant sour- ces. It is also well recognized that some con- taminants, such as Radon-222 gas and its radioactive decay daughters, and nitrogen di- oxide undergo chemical or radiochemical de- composition, and modelling the dispersal of these contaminants demands the additional consideration of the chemical or radiochemi- cal mass transformation processes involved.

It is also known that adsorption and de- sorption processes can affect the dispersal of contaminants in buildings. The evidence is pervasive. The lingering odours of winter clothes stored during the summer with naphthalene crystals, and the odour of clothes exposed to tobacco smoke or gasoline fumes that persists for days until washed are but two examples. Few would doubt that ad- sorption and desorption phenomena can and do play a role in the dispersal of airborne contaminants in buildings but, from the point of view of indoor air quality analysis, should we suspect these phenomena to be significant? Can adsorption and desorption significantly alter the dynamics of the disper- sal of indoor air contaminants? If so, under what circumstances will the effect be signifi- cant and how may one model the dispersal of

Page 2: Adsorption Modelling for Building Contaminant Dispersal Analysis

148 Axley: Adsorption Modelling

a contaminant when these circumstances oc- cur?

This paper demonstrates that adsorption and desorption phenomena have not only the potential to significantly alter the dyna- mics of contaminant dispersal but may, in some cases, completely control them. Work- ing from fundamental principles of adsorp- tion science, two families of adsorption/de- sorption models are formulated that provide a means to identify the circumstances under which sorption phenomena may come to be significant.

Fundamental Considerations A considerable body of knowledge is avail- able in existing adsorption science and tech- nology literature. Most of this literature is directed toward industrial applications of ad- sorption that involve the purposeful use of specially prepared adsorbents (e.g., activated carbon, zeolites, and synthetic resins) placed in adsorption devices that are carefully de- signed and controlled to achieve high rates of adsorption (e.g., well-mixed batch reac- tors, well-mixed flow reactors, and fixed-bed reactors) Uaroniec and Madey, 1988; Oscik and Cooper, 1982; Ponec et al., 1974; Rodri- gues et al., 1988; Ruthven, 1984; Slejko, 1985). Our interest here, on the other hand, is dir- ected toward modelling adsorption pheno- mena that are largely unintentional, invol- ving materials that tend to act as adsorbents, but are not specially prepared to do so, that are placed in buildings for purposes other than that of adsorption separation. Conse- quently, we shall have to adapt ideas from the available adsorption literature and limit consideration to the simpler models of ad- sorption phenomena until additional study warrants the application and provides the re- quisite data for more complete adsorption models.

Adsorption Dynamics Adsorption involves the separation of a sub- stance from one phase, indoor air in the

present context, and the accumulation of that substance on the surface of another phase, here, building materials. The nature of ad- sorption will depend on the nature of the ad- sorbent, the nature of the adsorbate, and the interaction that may occur between them. According to Slejko (1985), we may dis- tinguish four generic types of adsorption - exchange, physical, chemical, and specific adsorption. Exchange adsorption involves the exchange of one ionic species - the ionic ad- sorbate - with other ionic species attached to the surface of the adsorbent and therefore in- volves electrostatic attraction of the adsor- bate to the adsorbent. In physical adsorption the adsorbate is bound to the adsorbent by relatively weak intermolecular Van Der Waals forces. Chemical adsorption or chemi- sorption results from chemical reactions be- tween the adsorbate and the adsorbent, and therefore involves usually very strong bind- ing forces associated with the sharing of elec- trons between the adsorbate and adsorbent. Finally, the term specijic adsorption is reserved for adsorption phenomena involving binding of molecular groups, without chemical trans- formation, to specific functional groups on the adsorbent (e.g., polar adsorbates binding to polarized sites on the surface of an adsor- bent).

The nature of the adsorption process is directly linked to the strength of the forces that bind adsorbate to adsorbent with physi- cal adsorption falling at the weak-force end of the spectrum and chemical adsorption at the strong-force end. With this in mind, it follows directly that:

(a) adsorption is typically an exothermic process with the heat of adsorption asso- ciated with physical adsorption expec- ted to be small relative to that associated with chemical adsorption;

(b) physical adsorption may be expected to be a relatively reversible process while chemical adsorption may not; and

Page 3: Adsorption Modelling for Building Contaminant Dispersal Analysis

Axley: Adsorption Modelling 14Y

(c) chemical adsorption may be expected to be associated with the formation of a single layer of molecules of adsorbate on the surface of the adsorbent while in physical adsorption the possibility of multiple layers is more likely.

Reversibility and the heat of adsorption of a given adsorbate/adsorbent pair thus provide a direct indication of the molecular mechan- ism responsible for the adsorption.

Physical and some chemical adsorption processes may be more or less reversible and can be represented as:

Adsorbate + Adsorbent S

Adsorbate-Adsorbent Complex

where the degree of reversibility depends, in part, on the thermodynamic state of the ad- sorbate-adsorbent system and the nature of bonding in the adsorbate-adsorbent complex. For physical adsorption the bonding is rela- tively weak and thus this reversible process may be thought, instead, to be a reversible exchange between two alternative phases of the adsorbate - the free or gas phase and the adsorbed phase - where the adsorbent may be thought to activate the phase transforma- tion, not unlike a catalyst, rather than reac- ting to form a new adsorbate-adsorbent com- plex. Distinguishing the concentration of ad- sorbate near a surface, C*, from that ad- sorbed on the surface, Cs, this reversible physical adsorption process may be repre- sented as:

adsorbent c* 8 Cs + AH

where AH is the heat of adsorption and, for our purposes, concentration will be expres- sed in terms of mass fraction (i.e., for C* mass adsorbate per mass gas phase and for Cs mass adsorbate per mass adsorbent1).

The adsorption process is but one step of several that determine the nature of the ad- sorption dynamics. In the building context, adsorbate species are transported via diffu-

sion transport processes from the bulk air phase in building zones to near-surface loca- tions where the adsorption processes come into play to bind the adsorbate to the adsor- bent surface. Simultaneously, desorption pro- cesses release adsorbate from the adsorbent surfaces to near-surface locations where, now, diffusion transport processes act to transport the adsorbate to the bulk air phase. The rate of species transport to and from the adsor- bent surface is thus determined by both the rate of diffusion transport processes and ad- sorption-desorption kinetics.

Diffusion Transport Diffusion transport to and within effective adsorbents invariably involves a complex variety of transport mechanisms including boundary layer molecular and turbulent dif- fusion from the bulk air phase and molecu- lar, Knudsen, and surface diffusion within macroporous and microporous interstices (Ruthven, 1989). For our present purposes it will be sufficient to consider only boundary layer diffusion which may be modelled as a spatially discrete process using boundary layer theory (White, 1988); macroporous and microporous diffusion processes which are, by their nature, spatially continuous may be considered in future investigations using dis- crete Finite Element approximations. The pertinent details of boundary layer theory will be considered subsequently.

' The concentration of absorbate is sometimes expres- sed in terms of mass adsorbate per unit surface area available for adsorption or, in the indoor air quality literature, in terms of mass adsorbate per unit of pro- jected surface area of the adsorbent (e.g., wall surface area). Effective adsorbents typically have extremely large surface areas available for adsorption due to con- voluted surfaces and high porosities. As this surface area is well correlated to mass, for effective adsor- bents, it is usual to classify adsorbents by specific sur- face area (aredmass) and, hence, to employ mass frac- tion as a measure of concentration (e.g., see Vermeu- len et al., 1984).

Page 4: Adsorption Modelling for Building Contaminant Dispersal Analysis

150 Axley: Adsorption Modelling

Adsorption/Desorption Equilibria In typical situations the kinetics of the ad- sorption or desorption steps may be expected to be practically instantaneous relative to the diffusion transport processes and/or the rate of mass transport by advection into a given building zone; thus, it may not be necessary to explicitly model these kinetics. Although practically instantaneous, adsorption and de- sorption are limited by an equilibrium state. That is to say, for closed systems under stea- dy conditions the rate at which adsorbate molecules (or ions) bind to adsorbent sur- faces will eventually equal the rate at which they are released from the surface and the concentration in the free and adsorbed phases will remain constant at their respec- tive equilibrium values, C: and Cs,. In func- tional notation, we may say that the ad- sorbed phase equilibrium concentration is related to the free phase equilibrium concen- tration and the thermodynamic state of the system defined by temperature, T, and gas phase pressure, P, as:

These equilibrium relations are generally unique to the adsorbate-adsorbent system being considered and, when reported for iso- thermal conditions at atmospheric pressure, are identified as adsorption isotherm. Experi- mentally determined adsorption isotherms, typically represented graphically, may be ap- proximated by one of several equilibrium models including the Linear Model, Lang- muir Model, BET Model, and Freundlich Model - the first three having theoretical bases and the last being empirical.

ven, 1984; Slejko, 1985). This single par- ameter (i.e., K,) model may be expected to be appropriate for cases of physical adsorp- tion of gases at very low concentrations (e.g., trace amounts) on homogeneous surfaces with ample sites available for adsorption.

Lungmuir Model: In the years 1916-1918 Lang- muir developed an adsorption isotherm model that accounted for the fact that the sites available for adsorption are limited (Ruthven, 1984; Slejko, 1985). The Langmuir model may be expressed, for our purposes, as :

(3)

where Cso is the surface concentration corre- sponding to complete coverage by a single monolayer (constant, for a given adsorbate- adsorbent system, and independent of tem- perature for the Langmuir assumption of a fixed number of sites available for adsorp- tion) and KL is the temperature-dependent Langmuir adsorption coejjicicient. The Lang- muir model approximates the linear model for very low concentrations of adsorbate in the gas (or liquid) phase (i.e., when KL C: < < 1); thus we should expect near equality between the partition coefficient and the pro- duct of the parameters of this model as:

(4)

BET Model: Brunauer, Emmett, and Teller developed a model that accounted for the possibility of multilayer adsorption that may be expected to occur in physical adsorption (Ruthven, 1984):

Linear Model: This adsorption isotherm is defined by a simple linear relationship:

Cs, KpC, (2)

where K, is a temperature-dependent coefi- cient known as the partition coeficient (Ruth-

where KBET is the BET constant related to the energy of adsorption, Cso is the surface

Page 5: Adsorption Modelling for Building Contaminant Dispersal Analysis

Axley: Adsorption Modelling 151

concentration corresponding to complete coverage by a single monolayer (constant for a given adsorbate-adsorbent system), C’,,, is the air phase concentration corresponding to saturated conditions, and CX = (C’$C’,,,) is the reduced equilibrium concentration. The BET model has been found to be reliable in the range of reduced concentrations of 0.05 < (C7C’,J < 0.35 (Ruthven, 1984) for many gas adsorbate-solid adsorbent systems.

The BET model also approximates the linear model for very low concentrations of adsorbate in the gas (or liquid) phase:

thus we should expect near equality between the partition coefficient and the product of the parameters of this model as:

( Cso KBET/~* ) Kp = sat

Freundlich Model: An alternative two-par- ameter model that often fits experimental data well was promoted by Freundlich (Slej- ko, 1985):

where KF and n are the constant parameters associated with the Freundlich model.

From a somewhat simplified considera- tion of adsorption thermodynamics it may be shown that the temperature dependency of the adsorption coefficients used in the Linear, Langmuir, and BET models may be expected to be described by Arrhenius type relationships of the general form K = KgAwRT where K, is a constant, AH is an energy change associated with the adsorption pro- cess, R is the universal gas constant, and T is absolute temperature (Ruthven, 1984; Slejko, 1985). A plot of the logarithm of experiment- ally determined model constants log(K,), log(KL), or log (KBET) versus VT thus pro- vides one indication of the correctness of the

particular model (i.e., it should be a straight line) and provides a means to evaluate the Arrhenius constants. Borrazzo recently re- ported this to be so for some nine hydrocar- bons commonly found in indoor air ad- sorbed on nylon, wool, and glass fibres (An- delman, et al., 1989).

Building Zone Adsorption Dynamics Models The foregoing principles of adsorption sci- ence may be directly applied to the problem of modelling adsorption dynamics in build- ings. We begin by considering the adsorption dynamics of a single contaminant species cx in a well-mixed building zone containing a volume of air with mass m, and a quantity of adsorbent material with mass ms (see Figure 1). The (average) concentration of the conta- minant species in the zone air will be identi- fied as “Cz and the average concentration in the adsorbent as CS, both expressed in terms of mass fraction (i.e.,(mass a)/(mass air) and (mass a)/(mass adsorbent) respectively).

If the mass flow rate of air into the zone is w and the species concentration flowing into the zone is “C,,, then the mass flow rates of species cx into the zone and out of the zone “W,,, and “W,,,, are simply equal to the pro- duct of the respective concentrations and the air mass flow rate as:

(9)

In an analogous manner we may describe the species mass flow rate by adsorption and de- sorption transport between the room and the adsorbent material in terms of discrete spe- cies mass flow rates, (IWads and UWdes, that will, in general, depend on the zone and ad- sorbent concentrations as well as the physical characteristics of the adsorbate species a, ad-

Page 6: Adsorption Modelling for Building Contaminant Dispersal Analysis

152 Axlev: AdsorDtion Modellina

Adsorbent Material aCs 9 ms

aWO“t= aCZW Building Zone I‘ aCz 9 mz

~

Fig. 1 Zone plus adsorbent control volume

Fig. 2 Zone control volume and adsorbent control volume.

sorbent material, and the detailed nature of the airflow in the room:

“Wads = fads (.C, “CS, ... )

“wdes = fdes (“CZ, “CS, ...

(10)

(W

Isolating the control volume surrounding the adsorbent from the control volume sur- rounding the zone air, the adsorption and de- sorption mass flow rates may be diagramma- tically represented as shown in Figure 2.

Limiting consideration for the moment to completely reversible adsorption (i.e., allow- ing no irreversible destruction or transfor- mation of the adsorbate species a in the ad-

sorbent material or in the building zone) we may immediately write species mass balance relations for the adsorbent material, the building zone, and the systems as a whole as:

Building Zone Mass Balance

(“wout - “win) -k (awads - “Wdes) +

d“C, m, - = “Gz dt

Adsorbent Material Mass Balance

Page 7: Adsorption Modelling for Building Contaminant Dispersal Analysis

Axley: Adsorption Modelling 153 -

Building Zone and Adsorbent Material Mass Balance

daCs + m, - d%, ( a ~ o u t - “Win) + mz - dt dt

= “G, + aG, (14)

where we admit the possibility of generation of the contaminant species CI in both the building zone, “GZ and the adsorbent ma- terial “Gs.

Irreversible destruction or transformation of the contaminant species in the adsorbent or in the building zone may be introduced by inserting appropriate transformation ki- netics expressions for either or both of the generation terms above, *GZ and “Gs. Thus, for example, if the analyst wishes to model linear destruction kinetics in the adsorbent, “Gs would be replaced by an expression of the form “Gs = (-msKS) uCs, where tcS is the first order rate constant associated with the destruction. To allow a focused consideration of (reversible) physical adsorption, irrevers- ible transformation will not be considered further.

Equilibrium Adsorption To provide some insight into the dynamics of this coupled system it is useful to consider a limiting case where the adsorbed species concentration remains at all times in equi- librium with the zone concentration (i.e., diffusion, adsorption, and desorption trans- port rates are practically instantaneous rela- tive to the flow transport dynamics). More specifically, if the equilibrium relationship is described by the Linear Model:

“Cs = K p W Z (154

or:

dUCs - -- - dt

the mass Equation fies to:

daC, KP - dt

balance for the system as a whole, 14 with Equations 8 and 9, simpli-

Linear Equiltbrzum Adsorption Model Wac, + (m, + Kpm,) daC, __ = CIE (16a)

(16b)

dt

a E = (aG, + aG, + WaCi,)

where all forcing terms have been collected on the right-hand side to define the system excitation “E.

Alternatively, if the equilibrium relation- ship is described by the Langmuir Model:

or:

the mass balance for the system as a whole simplifies to:

Langmuir Equilibrium Adsorption Model

Wac, + (m,+ ( (1 + KL~C,)’ dt

For both cases the adsorbent acts to increase the participating mass of the building zone from that of the mass of the air in the zone, mZ, to this quantity plus the mass of the ad- sorbent, scaled by a factor. For the Linear Model, the effective participating mass is:

For the Langmuir Model the scaling factor includes a nonlinear dependency on the zone concentration:

but for situations where the zone concentra- tion does not vary greatly, this factor will re- main reasonably constant as well. Further-

Page 8: Adsorption Modelling for Building Contaminant Dispersal Analysis

154 Axley: Adsorption Modelling

more, for those situations where the zone concentration is small enough so that the KL"CZ term in the denominator may be ne- glected, then this factor should be expected to equal the partition coefficient of the Line- ar Model (as discussed earlier).

Dynamic Character of the Equilihum Adsorption Model For a single well-mixed zone under condi- tions of constant flow the system time con- stant, T, (or its inverse, the nominal air change rate);

T = (mzlw) ; w = constant (21)

provides a concise and convenient measure of the system's dispersal dynamics. For a well-mixed zone under conditions of con- stant flow with equilibrium adsorption dyna- mics added it follows that the (instantan- eous) system time constant, 7, becomes:

T = (meffective/w) ; w = constant (22)

Adsorption thus has the effect of increasing the system time constant in proportion to the adsorbent mass.

Example Application: Equilibrium A d s q t w n in a Room To fix these ideas and to evaluate the poten- tial significance of the participating mass of- fered by real adsorbents in building con- struction it is usehl, at this point, to con- sider a practical example. To this end, con- sider a room 5 x 8 m in plan with a 3 m floor to ceiling height, 25 mm thick nylon carpet- ing, and 13 mm thick gypsum board walls and ceiling.

The approximate weight of the gypsum board may be estimated at 10 kg/m2 and that of the nylon fibres in the carpeting at 2.5 kg/ m'. Thus, given the room geometry, the total mass of gypsum, mgyp, board is approximate- ly 11.8 x lo5 g and the mass of the nylon fibres in the carpet, mnylon, approximately 1.00 x 105 g.

At standard conditions of temperature and pressure the density of the air in the room is very nearly 1.2 kg/m3; thus the mass of air in the room, m,, is 1.44 x lo5 g.

Researchers at the Oak Ridge National Laboratory report a partition coefficient for the formaldehyde (HCH0)-gypsum wall- board system at 23 "C and 50% RH to be ap- proximately 5.5 g-air/g-gypboard (Matthews et al., 1987). (Silberstein also determined the same value for the partition coefficient but did not publish this result (Silberstein, 1988; Silberstein, 1989b).) If we assume that equi- librium adsorption conditions prevail, then for this measured partition coefficient and the representative room considered here, the gypboard may be expected to contribute a participating mass of HCHo-gFK P m gyp = 5.5 (g-air/g-gypboard) x 11.8 x lo5 (g-gypboard) = 65 x lo5 (g-air) that increases the participa- ting mass to 46 times the room air mass alone. This increase in participating mass would result in an equal increase in the time constant of the well-mixed zone, for condi- tions of constant airflow. Thus, for example, for a well-mixed zone without adsorption having a time constant of 1 hour (i.e., an air exchange rate of 1 air change per hour) the gypboard would alter the time constant for HCHO dispersal to 46 hours.

In another study (Andelman et al., 1989) partition coefficients for nylon, wool, and glass fibres are reported for some nine differ- ent aromatic and chlorinated hydrocarbons, designated simply as H-C-C1 here. For nylon fibres these partition coefficients range from 2.4 to 16.8 x lo4 g-air/g-nylon fibres.'

If, again, we assume that equilibrium ad- sorption conditions prevail, then the nylon fibres in the carpet of the representative room being considered may be expected to contri-

' The reported partition coefficient of 0.2 to 1.4 (cm3- air/g-nylon) was convened to the mass fraction units of (g-air/g-gypboard) based on an air density of 1.2 kg/m3.

Page 9: Adsorption Modelling for Building Contaminant Dispersal Analysis

Axley: Adsorption Modellins 155

bute a participating mass of HCC1-ny’onKpmnyton = (2.4 to 16.8) x (g-air/g-nylon) x 1.0 x lo5 (g-nylon) = (24 to 168) (g-air). In this case the room air mass would be effectively in- creased only slightly by 0.01 to 0.10%.

These simple calculations, based on the equilibrium adsorption assumption, indicate that adsorption can dramatically alter the dy- namics of contaminant dispersal in rooms when the product of the adsorbent mass and its partition coefficient are significant. Some initial experimental studies of adsorption in rooms support this conclusion (Matthews et al., 1987; Silberstein, 1989a).

Boundary layer Diffusion Controlled Adsorption In general, adsorption mass transport rates are determined by the individual transport rates of the diffusion processes and the sur- face adsorption kinetics that together ac- count for the multiple steps in the overall ad- sorption transport process. Throughout the adsorption literature, however, it is empha- sized that adsorption dynamics, in most practical situations, are invariably controlled by diffusion transport (the adsorption kine- tics step is practically instantaneous) although exceptions to this general rule exist (Ethier and Kamm, 1989). With this in mind, we will consider models for room ad- sorption dynamics that account for external film diffusion based on the assumptions that:

(a) equilibrium exists at the solid-gas inter- face at all times as defined by one of the equilibrium isotherms considered above (ie., the kinetics of adsorption is practic- ally instantaneous); and

(b) the system remains practically isothermal (i.e., the rate of heat generation due to adsorption is negligibly small and the heat transfer processes are sufficiently large to maintain isothermal conditions).

Models for room adsorption dynamics that account for macropore diffusion and micro-

pore diffusion will be considered in subse- quent studies.

Adsorption mass transport rates from the room air to solid (ie., nonporous) adsorbents with reasonably smooth surfaces may be ex- pected to be controlled by the rate at which a contaminant species diffuses from the bulk phase to the surface of the adsorbent. The diffusion mass transport rate will, in general, depend on the details of the contaminant concentration and velocity fields in the room, details that are both difficult to meas- ure and to determine analytically (e.g., by computational fluid dynamics) and, conse- quently, details that normally will be neither known nor easily determined. For wall sur- faces and other larger flat surfaces in rooms it is likely that any flow that exists will be directed more or less parallel to these sur- faces, forming a boundary layer over the sur- face of interest. Therefore, to obtain esti- mates of diffusion mass transport for these surfaces, we may reasonably apply some key results from boundary layer theory.

The adsorbent surface may be thought to be separated from the room air by afilm of air (i.e., the boundary layer) over which the contaminant concentration varies from a near-surface, air-phase concentration, CZ, to the bulk air-phase concentration in the zone, ’XZ, as illustrated in Figure 3.

If mass transport to the adsorbent is due only to film diffusion (e.g., there is no bulk flow of Contaminant through the adsorbent due to pressure gradients within or across the adsorbent) and the air-phase concentra- tions are not varying too rapidly, then the net mass transport rate from the bulk phase to the surface, uWs, may be approximated using steady-state mass transfer relations from boundary layer theory (White, 1988) of the general form:

Page 10: Adsorption Modelling for Building Contaminant Dispersal Analysis

156 Axley: Adsorption Modelling

3 aC;

aCz

Fig. 3 Boundary layer diffusion controlled adsorption

where: - hln

P

As

a - airD

ReL < 500,000 (244 is the average film mass transfer coeffi- cient acting over the adsorbent surface (lengthhime) - is the film density of air, the average of hm

a - airD the bulk and surface densities ReL > 500,000 ; SC = 1.0 is the projected surface area of the adsor-

bent

S ~ L E = 0.037 ReL W k Y 3 ; for

(24b)

The average film mass transfer coefficient may be measured directly (e.g., see Matthews et al., 1987) or indirectly using the naphtha- lene sublimation technique (White, 1988). It may also be estimated from published heat transfer coefficients or correlations (e.g., see Khalifa and Marshall, 1990) using the so- called heat and mass transjir analogy. Finally, several published mass transfer correlations are available that relate a dimensionless form of the average jilm mass transport coefficient, the average Sherwood number, S % , with the surface flow Reynolds number, ReL, and the air phase Schmidt number, Sc (White, 1988) (see also the closely related correlations for deposition velocity discussed by Nazaroff (Nazaroff and Cass, 1987; Nazaroff and Cass, 1989)). For flow parallel to a flat plate, a flow condition that may be considered to be rep- resentative of airflow past interior building surfaces, White provides the following corre- lations;

where:

UL ReL = __ 1,

U is the mean fluid velocity (parallel to the surface) outside of the boun- dary layer is the length of the surface in the direction of the flow

a-airD is the kinematic viscosity of the air phase

is the molecular difisivity of the binary a-air system being con- sidered.

L

v s c = ~

1,

a-aim

For many gases in air the molecular diffusiv- ity is practically constant with concentration (although it varies directly with temperature and inversely with pressure) with values, at room temperature aild atmospheric pressure,

Page 11: Adsorption Modelling for Building Contaminant Dispersal Analysis

Axley: Adsorption Modelling 157 -

ranging from to m2/s and the Schmidt number is typically close to 1.0 (e.g., 0.2 to 2.0).

With the expression for film diffusion transport, Equation 23, in hand we may for- mulate equations that describe the system adsorption dynamics by first demanding equilibrium at the adsorbent surface using either the Linear Model:

@Cs = KpaCi (25)

or the Langmuir Model:

‘ycs = aCs,,KLa~d (26) 1 + K L ~ c ~

and then substituting these equations and Equations 8, 9, and 23, into the basic mass balance relations for the system, Equations 12 and 13, to obtain, after simplification:

(a) for the Linear Model:

+ m Z - d a C ~ = ‘YG, (27a) dt

= aGs or, alternatively:

Boundary Layer Dffusion Controlled Ads* tion Model

where we have collected all forcing terms on the right-hand side as the system excitation vector {aE):

(b) for the Langmuir Model:

mS dorC, = a~~ dt

or, alternatively:

Langmuir Bounahry Layer Diffusion Controlled Adsorption Model -

dt = {aE} ( 2 8 ~ )

(The Langmuir Model Equations 28 are va- lid only up to the condition of complete cov- erage of the surface (i.e., (Cso - C,) > 0:) at which point Equations 28 become singular.)

Concise Matrix Notation The Linear Model and Langmuir Model variants of the system adsorption dynamics

Page 12: Adsorption Modelling for Building Contaminant Dispersal Analysis

158 Axley: Adsorption Modelling

equations considered so far may be repre- sented in concise notation as:

[“W]{aC) + [“MI = {aE} (29a)

where:

(a) for Equilibrium Adsorption:

dt

{“C} = “CZ

[“W] = w

Linear Model

[aM] = mz + KpmS (294

or

Lungmuir Model

[“MI = m, + ( aCsoKL ) ms (29e) (1 + KL“CZY

(b) for Boundaty Luyer D z f i i o n Controlled Adsorptwn

Linear Model

(290

or

Lungmuir Model

Note that the Langmuir Model introduces nonlinearity in the mass matrix, [“MI, of the Equilibrium Adsorption (EA) Model (i.e., with respect to zone concentration, “C,) and in the transport matrix, [“W], of the Boundary Layer Diffusion Controlled Adsorption (BLDC) Model (i.e., with respect to zone concentration, “Cs).

Dynamic Character of the BLDC Adsorption Model The essential dynamic character of the (sca- lar) EA Model is contained in the system time constant, t, or equivalently the system eigenvalue, A = Vt which is equal to A =

pM]- ‘[“W] = (w/meffective). Analogously, the essential dynamic character of the coupled BLDC Model equations is contained in the coupled system eigenvalues that satisfy the characteristic equation of the system:

det I [“MI’ [“W] - h [I] I = 0 (30)

where, for the Langmuir Model, this eigen- value problem would be evaluated at each specific state of the adsorbent concentration, OLCS, given the nonlinearity of the mass trans- port matrix in this case. In general, this ei- genvalue problem has two solutions, {Al, A2); the inverses of the system time constants, {t,, t2}. For the Linear Model, by substituting Equations 29g 8z h, into Equation 30 we ob- tain two solutions of a quadratic relation that correspond to these eigenvalues:

I -

(29i)

Page 13: Adsorption Modelling for Building Contaminant Dispersal Analysis

Axley: AdsorDtion Modellina 159

It is interesting to consider three limiting cases of this eigenvalue solution:

No D f i s i o n : When diffusion from the building zone to the adsorbent surface drops to zero, “h,pA, = 0, the system eigenvalues correspond to the eigenvalue of the well- mixed zone without adsorption as expected:

-

when: hm p As = 0 (32)

No Infiltratwn: When the infiltration into the zone drops to zero, W = 0, the system eigen- values become :

[E2] =

0

when: w = 0 (33)

Note: the first eigenvalue increases with in- creasing “h,pAs with the result that the sys- tem time constant, T~ = l/Al, approaches zero as the mass transfer rate increases.

Equilibnum Adsorption: If the rate of diffi- sion, “h,pAs, is large relative to the infiltra- tion mass flow rate, w, then the system eigen- values approach those corresponding to the Linear Equilibrium Adsorption Model:

[ 2 ; 2 ] -

when: h< p As > > w (34)

Again the first eigenvalue increases linearly with diffusion mass transfer rate. (The proof of Equation 34 is rather indirect. The sums (“FmpAs + w) in Equation 31 may be re- placed by “&,PA, when “h,pA, > > w and the resulting equation simplified and expanded using a Taylor’s series approxima- tion of the square root term to obtain the final result.)

Example Application: Boundary Layer D i h w n Controlled Adsorption in a Room To fix these ideas and to evaluate the signifi- cance of the diffusion transport, reconsider the room studied earlier using the Adsorp- tion Equilibrium Model, now applying, in- stead, the BLDC Model. This time we will limit consideration to the adsorption of for- maldehyde, HCHO, by gypsum board having the partition coefficient considered earlier, KP = 5.5 g-air/g-gypboard. The area of the adsorbent in this case is As = 118 m2. Assuming air flows uniformly along the length of the room, a reasonable estimate of the effective length of the adsorbent surface would be L = 8 m and an estimate of the free stream velocity of the airflow would be U = 8 m/h = 0.00222 m/s (i.e., given the air change rate of 1 air change per hour, the volumetric flow divided by the room cross- sectional area). Thus the effective Reynolds number would be (with the kinematic vis- cosity of the air of v = 1.5 x m2/s):

= 1,173 ReL = __ UL = (0.0022)(8)

V 1.5 x

For a Schmidt number of 1.0 we thus obtain the average Sherwood number for this prob- lem, from Equation 24a, as:

~

ShL = 0.664 ReL y2Scfi =

0.664(1173)fi (l.O)v3 = 22.7

For a representative gas diffisivity of HCHO-

”“D = 1.4 x m2/s we obtain an estimate of

Page 14: Adsorption Modelling for Building Contaminant Dispersal Analysis

160 Axley: Adsorption Modelling

the film diffusion mass transfer coefficient from Equations 24 as:

h,=sh~( HCHO-aia ) = 2 2 . 7 ( 1.4 ) = 3.97 x m/s

With these values in hand, an infiltration rate of W = 40 g/s (i.e., based on 1.0 air change per hour and an air density of 1,200 g/m3), and the zone and adsorbent masses computed earlier - mZ = 1.44 x lo5 g and mgyp = 7.8 x lo5 g - we may apply Equation 31 to compute the eigenvalues for the gyp- boardroom system:

kl = 3.17 x

A2 = 1.15 x

s-l = 1.14 hr-’ or tl = 0.876 hr

s-l = 4.14 x h i ’ or t2

= 242 hr

revealing a bimodal response characterized by a very long time constant of 242 h (i.e., 10.1 days) and a quick response somewhat smaller

than the nominal system time constant with- out adsorption (i.e., (mzN) = 1.0 h). This sort of response behaviour has, in fact, been observed in experimental studies (Dunn, 1987; Dunn and Tichenor, 1987; Matthews et al., 1987; Seifert and Schmahl, 1987; Silber- stein, 1989a); the results of these studies will be considered subsequently. The response of this system, and other systems governed by the BLDC Model, is particularly sensitive to boundary layer diffusion rate as character- ized by the average Sherwood number. Plot- ted below, for this example case, is the varia- tion of the system time constants for a varia- tion of average Sherwood number.

It is seen that the first time constant, 71 di- minishes from 1.0 h to zero as the second time constant, T~, drops from very high va- lues for an average Sherwood number near zero to an asymtoptic value for large values. In the present case this asymtoptic value is 30.8 hours and is equal to the time constant we obtained for this case using the Equilibri- um Adsorption Model (i.e., as predicted by Equation 34).

1

71 (h)

0.8

0.6

0.4

0.2

0

0 1000 2000 3000 4000 ShL 5000

Fig. 4 Time constont variation for the exomple application.

400

320

240

160

80

0

Page 15: Adsorption Modelling for Building Contaminant Dispersal Analysis

Axley: Adsorption Modelling 161

We may conclude, then, that at one ex- treme, with a near-zero rate of diffusion mass transfer, the system will respond to changes of concentration in the zone as a simple well-mixed zone with a time constant nearly equal to (mZrJu). For nonzero rates of diffu- sion the system will rapidly respond to changes of concentration with a time con- stant somewhat less than the simple well- mixed zone time constant; then, if the sec- ond mode is excited (e.g., if the contaminant is generated or concentrated in the adsor- bent) a very slowly changing response will be observed. Diffusion in smaller macropores and/or diffusion in micropores (e.g., having pore diameters approaching molecular sizes) may be expected to have, in essence, a simi- lar effect to that produced by very low mass transfer rates in the BLDC Model and may be responsible for this form of observed be- haviour. At the other extreme, very high rates of film diffusion mass transfer lead to the simple behaviour predicted by the EA Model as expected.

Surjbce Smoothness The BLDC Model presumes diffusion from the zone air to a smooth surface. This begs the question: what is meant by smooth? In the present context, it would seem reasonable to define an adsorbent surface as being “smooth” (i.e., sufficiently smooth to apply the BLDC Model) if:

(a) the surface roughness is small relative to the thickness of the boundary layer; and

(b) the structure of the surface roughness is coarser than that considered to be mac- roporous (ie., the width of surface irre- gularities or indentations are much lar- ger than the mean free path of the dif- fusing contaminant species) so that the diffusion obeys Fick’s Law.

The thickness of the boundary layer, 6, may be estimated as (Beek and Muttzall, 1975):

L 6 == ShL

(35)

For the example considered above we obtain 6 = 8 m / 27.9 = 0.29 m which is certainly large relative to the (likely) physical rough- ness of the gypboard surface, thus justifying the application of the BLDC Model in this case.

Adsorption Element Equations for General Multi-Zone Analysis Contaminant dispersal analysis theory may be formulated using an element assembly ap- proach wherein equations approximating the behaviour of the building airflow system or subsystem, the system equations, are assem- bled from equations that describe contami- nant mass transport in discrete portions or elements of the system model. The details of this approach are presented elsewhere (Axley, 1989). In this section we shall simply de- scribe element equations that correspond to the adsorption models used above in the single zone case that may be used for multi- zone system models of arbitrary complexity.

In general, for the dispersal of species a, contaminant dispersal element equations have the following form:

where {aCe) is a vector of contaminant spe- cies a discrete concentration variables asso- ciated with the element “eyy, a subset of sys- tem discrete variables {%I, and {”$) is a vector of element derived contaminant spe- cies generation rates. The element matrices [“W] and [“me] are square transformation matrices known as the element species trans- port matrix and species mass matrices re- spectively.

The element equations corresponding to the models presented above may be derived directly from Equations 29 by simply sub- tracting the single zone mass and airflow transport terms from the system equations:

Page 16: Adsorption Modelling for Building Contaminant Dispersal Analysis

162 Axley: Adsorption Modelling

Equilihum A d s q t w n Element: a single node element, say element “e”, with an adsorbent mass of m: associated with a well-mixed zone “i” with :

{ w e } = {ac;} (374

= [OI (37b)

Linear Model

[amel = KpmeS[l] (374

or

Boundary hyer D i m w n Controlled Ads+ tion: a two node element, say element “e”, as- sociated with a well-mixed zone “i” and an adsorbent “j’’ with an adsorbent mass of mes:

Linear Model

[a*] = (%,A3

or

Langmuir Model

[a*] = ( E p AeS)

-1 1 -

KP

KP

1 -1 -

where all variables are defined as earlier (those variables with a following superscript “e” are specific to the adsorption element “e”). Again, note that the Langmuir Models introduce nonlinearities, now at the element level, and, for the boundary layer diffusion controlled case, are defined for adsorbent concentrations less than

Implementation Using Existing Multi-Zone Program The Linear Models for both cases above may be implemented using existing multi-zone contaminant dispersal analysis programs that provide only well-mixed zones and discrete flow paths. For equilibrium adsorption mod- elling the analyst need only increase the as- sociated well-mixed zone mass by K,m: to account for mass transport to the adsorbent.

The boundary layer diffusion controlled case requires a slightly indirect tactic. By re- defining the element concentration variables as:

the element arrays may be rewritten as:

Linear Model

[aw‘] = ( g p AeS) 1 -1 -1 1

These element equations have the form of a subassembly of a well-mixed zone linked to inflow and outflow flow elements; thus, these element equations may be added to existing multizone contaminant dispersal models by linking a pseudo-well-mixed zone

Page 17: Adsorption Modelling for Building Contaminant Dispersal Analysis

Axley: Adsorption Modelling 163

to the actual well-mixed zone containing the adsorbent by two pseudo-flow paths (i.e., flow elements) having equal and opposite flows. By setting the pseudo zone mass to K,m: and the pseudo-mass flows to (he, p ks), the behaviour of the adsorbent will be modelled although the computed concentra- tion response in the pseudo-well-mixed zone must be scaled by K, to obtain the species concentration in the adsorbent, q.

Application and Validation In this section we will compare results from the EA and BLDC models with the results of adsorption/desorption studies of formalde- hyde on gypsum wall board reported by re- searchers at Oak Ridge National Lab (ORNL) and the National Institute of Stan- dards and Technology (NIST). Each of these groups developed a semi-empirical model. Similar semi-empirical sorption models and closely related sourcehink models have been reported by others (Clausen et al., 1990; Dunn, 1987; Dunn and Tichenor, 1987; Ti- chenor er al., 1990; Tucker, 1988). In contrast, the EA and BLDC models examined here are based on the theory presented above. Be- fore considering the details of these studies two other studies deserve mention.

Cleary and Sonderegger, at Lawrence Berkeley Lab, and Thomas and Burch, at

NIST, developed models to predict the ad- sorption and desorption of water in building materials (Cleary and Sherman, 1984; Cleary and Sonderegger, 1984, Thomas and Burch, 1990). Both groups considered models simi- lar to the BLDC model developed above, although the NIST researchers presented this approach as a limiting case of a more general model that included diffusion within the porous adsorbent. They combined boun- dary layer solutions with equilibrium ad- sorption isotherms to form an adsorption dy- namics model. In comparisons of modelled response to measured response these models proved accurate.

ORNL Study Researchers at the Oak Ridge National Lab investigated the adsorption and desorption characteristics of gypsum wall board speci- mens in a small test chamber and demon- strated that an empirical three-parameter ex- ponential response model could be fitted to measured response data with reasonable ac- curacy. A diagram of the test set up is shown in Figure 5.

In this investigation six specimens of gyp- sum board (0.25m x 0.24m) were placed in one of a pair of dual test chambers (volumes of 0.20 m3) and air, at 23 "C and 50% RH, containing a controlled concentration of for- maldehyde, "C,,, was passed through this

GvDboard Chamber Vol = 0.20 23OC. 50%

m3 RH

I Reference Chamber

Vol = 0.20 m3 23"C, 50% RH

Fig. 5 ORNL Dual chamber test facility.

Page 18: Adsorption Modelling for Building Contaminant Dispersal Analysis

164 Axlev: Adsorotion Modellina

chamber and an identical reference chamber at constant and identical rates. The exit for- maldehyde concentrations from each of the chambers, “c,, and “creb were measured at periodical intervals and the ratio, R, of these concentrations, adjusted for initial condi- tions and normalized, was reported. Finally, three-parameter (i.e., A, Z, and 7) exponen- tial relationships of the following forms:

R = Z + (A-Z)(l-e-t/T) for adsorption (40)

R = (A-Z)e-dT for desorption (41)

were fitted to measured response data. It must be emphasized, to use the words of these researchers, that “Although exponen- tial sorption and desorption processes may represent the physical behavior of the experi- mental system, the precise form of the mod- els is empirically derived”. The gypsum wall board chamber was modelled using the Line- ar variants of the Equilibrium Adsorption (EA) and Boundary Layer Diffusion Con- trolled (BLDC) models, Equations 16 and 27 proposed above, and the reference chamber was modelled as a well-mixed zone, as shown in Figure 6. (For these and the following

~ ~ ~ ~ ~ ~ ~~~~

computational studies the density of the gyp- sum wall board was estimated to be 769,000 &m3, the molecular diffusivity of formalde- hyde in air at 23 “C was estimated to be 1.39 x m2/s using the Chapman-Enskog for- mula (Bird et al., 1960; Satterfield, 1970), and the kinematic viscosity of air was estimated to be 1.53 x m2/s.)

The response of these modelled chambers was computed, using CONTAM87 (Axley, 1988), for a twenty-eight-day adsorption test followed by a four-day desorption test. The results of these analyses are compared in Fig- ures 7 and 8 to the ORNL empirical model along with an indication of its uncertainty (i-e., the + err and - err curves) based on the reported standard errors of the model para- meters obtained from nonlinear regression analysis.

It is seen that the initial response appears to be well-captured by the EA model and the longer term response is better captured by the BLDC Model. An examination of the ORNL raw data (see Figure 3 of Matthews et al., 1987) reveals greater scatter, especially in the bend of these curves, than indicated by the error bounds shown above; thus a re- sponse falling between the EA and BLDC

c ref

llllll Illill

mz = 240 g meff = 20,106 g

C in Cin Cin Reference Chamber Gypboard Chamber Gypboard Chamber Well-Mixed Model EA Model BLDC Model

Fig. 6 Idealizations of ORNL test Chambers.

Page 19: Adsorption Modelling for Building Contaminant Dispersal Analysis

Axley: Adsorption Modelling 165

1.20

R

1 .oo

0.80

0.60

0.40

0.20

0.00

- - ORNL+err

Fig. 7

1 .oo

R'

0.80

0.60

0.40

0.20

0.00

28 24 Days 0 4 8 12 16 20

Comparison of modelled adsorption response and ORNL empirial model.

4 Days 0 1 2

Fig. 8 Comparison of modelled desorptian response and ORNL empirical model.

Page 20: Adsorption Modelling for Building Contaminant Dispersal Analysis

166 Axley: Adsorption Modelling

Models would be considered practically ac- curate. The BLDC model is sensitive to the value used for the average film mass transfer coefficient L. In the present case, this coeffi- cient was estimated using the correlations defined by Equations 24 which require an es- timate of the flow Reynolds number. The flow Reynolds number was based on an esti- mate of the mean flow velocity in the cham- ber equal to the volumetric flow rate divided by the square of the cube root of the cham- ber volume (i.e., an estimate of the chamber cross-sectional area). More detailed infor- mation of the chamber geometry could not be obtained. This resulted in an estimate of Re = 4.25, an extremely low value; it is cer- tain that the obstruction of the test speci- mens would have resulted in a larger esti- mate. The effect of higher Reynolds numbers (i.e., greater diffusion rates) would bring the BLDC Model curve closer to the EA Model curve in the early response and thus improve the accuracy of the modelling.

The desorption response falls between the EA and BLDC model predictions. Again, the accuracy of the estimate of the flow Rey- nolds number is critical; a larger and, pre- sumably more reasonable, Reynolds number would have improved the BLDC modelled results by “pulling” the BLDC curve closer to the EA curve.

The adsorption dynamics models presen- ted in this paper depend critically on the equilibrium adsorption isotherm model em- ployed, in the present case, on the partition coefficient K,. A partition coefficient of 5.5 (g-air/g-gypboard) was used for these studies and the NIST studies below. This is a value that was determined by both ORNL and NIST investigators using a rather intuitive method based on integrating the desorption response data that corresponds, in essence, to an integral form of the EA Model equations, Equations 16, or:

where one integrates response and excitation values over an arbitrary time interval, t1...t2, and evaluates the change of concentration over the same interval:

Solving this equation for the partition coeffi- cient:

K, = msAaCz mS

defines the method used by these investiga- tors to determine K, (although both groups neglected to consider the second term on the right-hand side, (mz/ms), which was practic- ally insignificant in both studies).

This determination of the partition coeffi- cient implicitly assumes linear equilibrium adsorption behaviour and therefore stacks the deck in favour of the Linear Equilibrium Adsorption Model. Silberstein, in the NIST study, also, however, used an alternate strat- egy to determine the partition coefficient and obtained practically identical results, thereby providing some validation of the value used here. It would be best to deter- mine this equilibrium adsorption parameter or, better, the equilibrium adsorption charac- teristics using one of several available inde- pendent means to do so (Ruthven, 1984).

NIST Study In tests to study formaldehyde levels in a test house due to emissions from a variety of pressed-wood products, researchers at the National Institute of Standards and Tech-

Page 21: Adsorption Modelling for Building Contaminant Dispersal Analysis

Axley: Adsorption Modelling 167 -

............................................... ........................................................... \ , , , 23OC \ , \

\ \

5 23°C 50%RH CZ

\

\

llllll I llllll m,u = 161,056 \

\ \

, \ \ , ,

: w=2201 g/ \

\ \ \ c=o .......................................................... c=o ..............................................

NIST Chamber NIST Chamber EA Model BLDC Model

Fig. 9 Idealization of the NIST test chamber.

nology found that it took days to reach equi- librium concentration levels and suspected that these long delays were due to adsorption and desorption (Grot et al., 1985). To invest- igate this phenomena more directly these re- searchers conducted a separate investigation in a “medium-sized” dynamic measuring chamber, formulated a semi-empirical model of the dynamics of the sorption processes, and numerically extracted values for the model parameters from measured data (Sil- berstein, 1989a). Unlike the ORNL study, the environmental chamber employed was near- ly room size, the gypsum board was attached to the walls of the chamber, as it would be in building construction, exposing a single side, and the proposed model was based, qualita- tively, on the likely physical mechanisms re- sponsible for adsorption. This model was not, however, directly formulated from fun- damental principles of adsorption science and therefore must be considered a semi-em- pirical model.

In the NIST study a 1.2 m by 2.4 m speci- men of 0.013 m thick gypsum wall board was secured to a wall of a “medium size” test chamber with inside dimensions of 1.22 m x 2.44 m x 0.61 m. This test chamber was placed in a closed environment maintained

at 23 “C and 50% RH, a constant fresh air- flow rate through the chamber of 1.01 air changes per hour was maintained, and, for adsorption studies, formaldehyde gas was generated within the chamber at a constant rate.

The NIST chamber was modelled using the Linear variants of the EA and BLDC models estimating the density of the gypsum wall board, the molecular diffusivity of the formaldehyde, and the kinematic viscosity of the air, as above, and employing the partition coefficient of 5.5 (g-air/g-gypboard). Dia- grams of these models are shown in Figure 9. The response of these models to an 11-day adsorption test, with formaldehyde generated within the chamber at a constant rate of 0.24 mgh, followed by a 10-day desorption test, without the generation of formaldehyde, were computed using CONTAM87. The re- sults of these computational analyses are compared to the measured data in Figure 10.

Regrettably, there was some confusion re- garding the actual formaldehyde generation rate and infiltration rate used during the test. Test data were obtained from two draft re- ports of the tests (Silberstein, 1988; Silber- stein, 1989b). In one of these reports an infil- tration rate of 1.01 ACH was reported but the

Page 22: Adsorption Modelling for Building Contaminant Dispersal Analysis

168 Axley: Adsorption Modelling

0.15

[HCHO] rng/rn3

0.10

0.05

0.00

I I I I I

20.00 l6.O0 Days 0.00 4.00 8.00 12.00

Fig. 10 Comparison of modelled response and NlST measured data. Formaldehyde was generated within the chamber from day 0 through day 10.

uncertainty associated with this value was great. The formaldehyde generation rate was reported to be 0.313 mg/h for the adsorption tests yet hypothetical steady-state concentra- tion levels for an empty chamber were repor- ted that indicated the actual generation rate to be 0.24 mg/h. Computations were based on an air exchange rate of 1.01 ACH and a generation rate of 0.24 mg/h. (Silberstein suggested, in a personal communication, that an air exchange rate closer to 1.3 ACH and a generation rate of 0.313 mg/h might be more appropriate. Since the ratios of these values are nearly identical to the values used, the computed results would be very similar.)

As in the ORNL case, the measured re- sponse very nearly falls between the EA Model and the BLDC Model responses. Again the BLDC response was especially sensitive to the estimation of the average mass transfer coefficient that was, in this case, estimated using the correlations with the flow Reynolds number reported above. In this case the Reynolds number was esti- mated to be 110, a relatively low value. In

Figure 10 the BLDC Model response is plot- ted for a ten-fold increase in Reynolds num- ber (i.e., a fl increase in diffusion mass transfer rate) and it may be seen that the modelled response in this case more closely follows the measured response. Although a somewhat higher Reynolds number would have led to excellent agreement, no reliable means are available to estimate the flow Rey- nolds number accurately, clearly a limitation of the use of these correlations.

Conclusion Two families of macroscopic adsorption models have been formulated, based on fun- damental principles of adsorption science and technology, that may be used for the purposes of indoor air quality analysis. The first family of adsorption models - the Equi- l i h u m Adsorption (EA) Models - are based upon assuming that zone air concentrations and adsorbent concentrations remain in equilibrium at all times, where the equilibri-

Page 23: Adsorption Modelling for Building Contaminant Dispersal Analysis

Axley: Adsorption Modelling 169

um condition may be defined by any num- ber of equilibrium isotherms. Two members of this family are explicitly discussed, the first based on the linear isotherm model and the second on the Langmuir isotherm model. The second family of adsorption models - th Boundmy Lyer Diffusion Con- trolled Adsorption (BLDC) Models - are based on the use of boundary layer theory to model the rate of mass transfer from the bulk phase through the boundary layer to the solid adsor- bent phase and the assumption that equi- librium exists at the solid-gas interface. Again this equilibrium may be modelled using any one of several adsorption isotherm models. Two members of this model are ex- plicitly discussed, one based on the linear ad- sorption isotherm model and the other on the Langmuir model, as before.

An analysis of the linear variants of the EA and BLDC models, presented in this paper, demonstrated that the EA model and the simple well-mixed zone model (ie., with- out adsorbent) are simply limiting cases of the BLDC model. This analysis provides for- mal proof of the physical consistency of the BLDC approach.

It is shown that the linear variants of both the EA and BLDC models lead directly to contaminant dispersal sorption elements that may be used directly with existing contami- nant dispersal analysis programs such as the CONTAM family of programs. These mod- els have the attractive feature that they de- pend on a single parameter, the partition coefficient, that may be measured using one of several well-documented methods (Ruth- ven, 1984). They must be expected, however, to be limited to the modelling of physical ad- sorption of trace contaminants because the linear adsorption isotherm upon which they are based is generally only valid for very low adsorbate concentrations. The BLDC cap- tures the biphasic or bimodal nature that has been observed in adsorption and desorption tests (Dunn, 1987; Dunn and Tichenor, 1987; Matthews et al., 1987; Seifert and Schmahl,

1987; Silberstein, 1989a) but it relies on an es- timation of an average film mass transfer coefficient, an uncertain problem that de- mands knowledge of the nature of airflow in the zone which may not be available. DiffL sion through a boundary layer must be ex- pected, however, to be an important step in practically all cases of adsorption and, short of complete solution of the microscopic equations governing the flow in a zone, the boundary layer solutions to this problem are the best available; the analogous solutions for boundary layer heat transfer have served the building energy simulation community well for some fifty years (Khalifa and Mar- shall, 1990).

Applications of the linear variants of the EA and BLDC models to formaldehyde ad- sorption and desorption tests were presented to provide some first validations of these simple models. In these studies measured re- sponse data, reported by NIST and ORNL researchers, were compared with modelled response and the agreement was good. These practical applications and the example appli- cations considered to clarify the develop- ment of the models clearly indicate the sig- nificant, in fact, overwhelming effect that simple physical adsorption and desorption transport processes may have on the disper- sal of contaminants in buildings. Further- more, physical adsorption must be expected to be pervasive depending not so much on special characteristics of adsorbate-adsorbent systems but simply on the availability of suf- ficient adsorbent surface. To date, sorption dynamics have seldom been considered in practical indoor air quality analysis and, im- portantly, in the closely related field of tracer gas analysis of building airflows. Both areas of analysis should be re-evaluated in light of the implications of the results reported here.

The success of the work reported by Cleary (Cleary and Sherman, 1984; Cleary and Sanderegger, 1984) and Thomas (Tho- mas and Burch, 1990) in modelling moisture sorption in building materials supports the

Page 24: Adsorption Modelling for Building Contaminant Dispersal Analysis

170 Axley: Adsorption Modelling

hope that the nonlinear variants (e.g., Lang- muir and BET adsorption isotherms) of the EA and BLDC models will prove effective in modelling physical sorption mass transport when adsorbate concentrations vary over wide ranges and may not be at trace levels (e.g., moisture variations in buildings in gen- eral), but this will have to be investigated.

Not presented in this paper are physical adsorption and desorption models based on existing knowledge of macroporous and mi- croporous diffusion processes. It is clear, however, that macroscopic models could be readily developed based on finite element ap- proximations to available governing differen- tial equations for these classes of diffusion (e.g., the so-called Glueckauf Model for the diffusive transport of an adsorbate down a cylindrical pore in a macroporous adsorbent (Holland and Liapis, 1983)). The numerical studies reported clearly indicate the potential importance of diffusion rate limitations -

diffusion resistance can lead to extremely long time constants in adsorption dynamics. It is entirely possible that the exponentially- diminishing “source” models reported by some investigators and the apparent irrevers- ible, but small, conversion of formaldehyde in gypsum wall board are both due to more deeply adsorbed (e.g., within macroporous or microporous structures) constituents. It seems likely that a consideration of these more rate-limiting diffusion steps in the ad- sorption process may lead to some rationally based physical “source” models.

Acknowledgments The research reported was supported by the U.S. National Institute of Standards and Technology (NIST) and the U.S. Department of Energy (DOE) and this paper was abstrac- ted from a report prepared for these agencies (Axley, 1990).

Page 25: Adsorption Modelling for Building Contaminant Dispersal Analysis

Axlev: Adsorption Modellinq 171

References Andelman, J. B., Giardino, N. J., Marshall, J., Esmen, N.

A., Borrazzo, J. E., Davidson, C. I., Small, M. and Wilkes, C. (1989) Exposure to Volatile Chemicals from Indoor Uses of Water, Las Vegas, Nevada, Air & Waste Management Association.

Axley, J. W. (1988) “Progress toward a general analytical method for predicting indoor air pollution in build- ings: indoor air quality modelling phase 111 report” (NBSIR 88-3814). US. DOC, NBS, Gaithersburg, MD.

Axley, J. W. (1989) “Multi-zone dispersal analysis by ele- ment assembly”, Building and Environment, 24 (2),

Axley, J. W. (1990) Adsorption Modelling for Macroscopic Contaminant fipersul Analysis, National Institute of Standards and Technology (NIST/GCR/ 90/573).

nomena, London, John Wiley & Sons. Bird, R. B., Stewart, W. E. and Lightfoot, E. N. (1960)

Transpa Phenomena, New York, John Wiley & Sons. Clausen, l? A,, Wolkoff, l? and Nielsen, I? A. (1990) Long

Term Emission of Volatile Organic Compoundsfrom Wa- terborne Paints in Environmental Chambers, Toronto, Canada, Canada Mortgage and Housing Corporation, Ottawa, pp. 557-562.

Cleary, l? and Sherman, M. (1984) S e a s m l Storage of Moisture in Roof Sheathing Lawrence Berkeley Labora- tory, Applied Science Division (LBL-17774).

Cleary, I? and Sonderegger, R. (1984) A Method to Predict the Hour by Hour Humdity Ratio of Anic Air, Law- rence Berkeley Laboratory, Applied Science Division

Dunn, J. E. (1987) “Models and statistical methods for gaseous emission testing of finite sources in well- mixed chambers”, Atmosphenc Environment, 21 (2),

Dunn, J. E. and Tichenor, B. A. (1987) Compensating for Wall Eflects in I A Q Chamber Tests by Mathematical Modelling, New York, Proceedings of the 80th Annual Meeting of the Air Pollution Control Association.

Ethier, C.R. and Kamm, R.D. (1989) “Mass transfer dur- ing rate-limited Langmuir adsorption in a pore”, Phy- stochemical Hydrodynamics, ll(2), 205-217.

Grot, R. A., Silberstein, S. and Ishiguro, K. (1985) “Vali- dation of models for predicting formaldehyde concen- trations in residences due to pressed wood products”: Phase I. Gaithersburg, MD: U.S. D.O.C., NBS, NEL, CBT.

Holland, C. D. and Liapis, A. I. (1983) Computer Methods for Solving Dynmnic Separation Problems, New York, McGraw-Hill.

Jaroniec, M. and Madey, R. (1988) Physical Adsorption on Heterogeneous Solds, Amsterdam, Elsevier.

Khalifa, A. J. N. and Marshall, R. H. (1990) “Validation of heat transfer coefficients on interior building sur- faces using a real-sized indoor test cell”, International Journal of Heat and Mass Transfer, 33 (lo), 2219-2236.

Matthews, T. G., Hawthorne, A. R. and Thompson, C.

113-130.

Beek, W. J. and Muttzall, K. M. K. (1975) T r a n ~ p ~ n Phe-

(LBL-17591).

425-430.

V. (1987) “Formaldehyde sorption and desorption characteristics of gypsum wallboard”, Environmental Science and Technology, 21 (7), 629-634.

Nazaroff, W. M. and Cass, G. R. (1987) Mass Transfer Aspects of Pollutant Removal at Indoor Surfaces, West Berlin, Institute for Water, Soil, and Air Hygiene, pp.

Nazaroff, W. M. and Cass, G. R. (1989) “Mathematical modelling of indoor aerosol dynamics”, Environmental Science and Technology, 23 (2), 157-166.

Oscik, J. and Cooper, I. L. (1982) Adsorption, New York, Halsted Press.

Ponec, V., Knor, Z. and Cerny, S. (1974) Adsorption on Solrds, Cleveland, CRC Press.

Rodrigues, A. E., LeVan, M. D. and Tondeur, D. (1988) Adsorption: Science and Technology, Dordrecht, Kluwer Academic Publishers.

Ruthven, D. M. (1984) Pnnciples of Adsorption and Ad- sorption Processes, New York, John Wiley & Sons.

Ruthven, D. M. (1989) “Adsorption kinetics”. In: Rodri- gues, A.E. et al. (eds), Adsorption Science and Tech- nology, Dordrecht, Kulwer Academic Publishers, pp.

Satterfield, C. N. (1970) Mass Transfer in Heterogeneous Catalysis, Cambridge, MA, MIT Press.

Seifert, B. and Schmahl, H. J. (1987) Quantification of Sorption Effects for Selected Organic Substances Present in Indoor Air, West Berlin, Institute for Water, Soil, and Air Hygiene, pp. 252-256.

Silberstein, S. (1988) “A gypsum wallboard formalde- hyde sorption model”, National Institute of Standards and Technology (DRAFTNISTIR).

Silberstein, S. (1989a) “A gypsum wallboard formalde- hyde sorption model”, National Institute of Standards and Technology (NISTIR:894028).

Silberstein, S. (1989b) “A gypsum wallboard formalde- hyde sorption model”, National Institute of Standards and Technology (DRARNISTIR).

Slejko, E L. (1985) Adsopnion Technology: A Step-by-step Approach to Process Evaluation and Application, New York, Marcel Dekker, Inc.

Thomas, W. C. and Burch, D. M. (1990) “Experimental validation of a mathematical model for predicting water vapor sorption at interior building surfaces”, A S H R A E Transactions, 96 (Pt. 1).

Tichenor, B.A., Guo, Z., Mason, M. and Dunn, J. (1990) Evaluation of Indoor Air Pollutant Sinks for Vapor P h e Organic Compounds, Toronto, Canada, Canada Mort- gage and Housing Corporation, Ottawa, pp. 623-628.

Tucker, G. (1988) Air Pollutants from Surface M a t e d s : Facton Influencing Emlsswns, and Predictive Models, Stockholm, Sweden, Swedish Council for Building Research, pp. 149-157.

Vermeulen, T., LeVan, M.D., Hiester, N.K. and Klein, G. (1984) “Adsorption and ion exchange”. In: Perry, R.H., Green, D.W. and Maloney, J.O. (eds), Penys Chemical Engineer’s Handbook, New York, McGraw- Hill, 6th ed.

White, E M. (1988) Heat and Mass Transfer, New York, Addison-Wesley Pub. Co.

244-248.

87-114.

Page 26: Adsorption Modelling for Building Contaminant Dispersal Analysis