emissions of 2,2,4-trimethyl-1,3-pentanediol monoisobutyrate (tmpd-mib) from latex paint: a critical...

30
This article was downloaded by: [North Dakota State University] On: 07 November 2014, At: 14:24 Publisher: Taylor & Francis Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK Critical Reviews in Environmental Science and Technology Publication details, including instructions for authors and subscription information: http://www.tandfonline.com/loi/best20 Emissions of 2,2,4-Trimethyl-1,3- Pentanediol Monoisobutyrate (TMPD-MIB) from Latex Paint: A Critical Review RICHARD L. CORSI a & CHI-CHI LIN b a Department of Civil, Architectural and Environmental Engineering , The University of Texas , Austin, Texas, USA b Department of Civil and Environmental Engineering , National University of Kaohsiung , Kaohsiung City, Taiwan Published online: 02 Dec 2009. To cite this article: RICHARD L. CORSI & CHI-CHI LIN (2009) Emissions of 2,2,4-Trimethyl-1,3- Pentanediol Monoisobutyrate (TMPD-MIB) from Latex Paint: A Critical Review, Critical Reviews in Environmental Science and Technology, 39:12, 1052-1080, DOI: 10.1080/10643380801977925 To link to this article: http://dx.doi.org/10.1080/10643380801977925 PLEASE SCROLL DOWN FOR ARTICLE Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and should be independently verified with primary sources of information. Taylor and Francis shall not be liable for any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of the Content. This article may be used for research, teaching, and private study purposes. Any substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://www.tandfonline.com/page/terms- and-conditions

Upload: chi-chi

Post on 13-Mar-2017

214 views

Category:

Documents


0 download

TRANSCRIPT

This article was downloaded by: [North Dakota State University]On: 07 November 2014, At: 14:24Publisher: Taylor & FrancisInforma Ltd Registered in England and Wales Registered Number: 1072954 Registeredoffice: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Critical Reviews in EnvironmentalScience and TechnologyPublication details, including instructions for authors andsubscription information:http://www.tandfonline.com/loi/best20

Emissions of 2,2,4-Trimethyl-1,3-Pentanediol Monoisobutyrate (TMPD-MIB)from Latex Paint: A Critical ReviewRICHARD L. CORSI a & CHI-CHI LIN ba Department of Civil, Architectural and Environmental Engineering ,The University of Texas , Austin, Texas, USAb Department of Civil and Environmental Engineering , NationalUniversity of Kaohsiung , Kaohsiung City, TaiwanPublished online: 02 Dec 2009.

To cite this article: RICHARD L. CORSI & CHI-CHI LIN (2009) Emissions of 2,2,4-Trimethyl-1,3-Pentanediol Monoisobutyrate (TMPD-MIB) from Latex Paint: A Critical Review, Critical Reviews inEnvironmental Science and Technology, 39:12, 1052-1080, DOI: 10.1080/10643380801977925

To link to this article: http://dx.doi.org/10.1080/10643380801977925

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the“Content”) contained in the publications on our platform. However, Taylor & Francis,our agents, and our licensors make no representations or warranties whatsoever as tothe accuracy, completeness, or suitability for any purpose of the Content. Any opinionsand views expressed in this publication are the opinions and views of the authors,and are not the views of or endorsed by Taylor & Francis. The accuracy of the Contentshould not be relied upon and should be independently verified with primary sourcesof information. Taylor and Francis shall not be liable for any losses, actions, claims,proceedings, demands, costs, expenses, damages, and other liabilities whatsoever orhowsoever caused arising directly or indirectly in connection with, in relation to or arisingout of the use of the Content.

This article may be used for research, teaching, and private study purposes. Anysubstantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &Conditions of access and use can be found at http://www.tandfonline.com/page/terms-and-conditions

Critical Reviews in Environmental Science and Technology, 39:1052–1080, 2009Copyright © Taylor & Francis Group, LLCISSN: 1064-3389 print / 1547-6537 onlineDOI: 10.1080/10643380801977925

Emissions of 2,2,4-Trimethyl-1,3-PentanediolMonoisobutyrate (TMPD-MIB) from Latex Paint:

A Critical Review

RICHARD L. CORSI1 and CHI-CHI LIN2

1Department of Civil, Architectural and Environmental Engineering, The University of Texas,Austin, Texas, USA

2Department of Civil and Environmental Engineering, National University of Kaohsiung,Kaohsiung City, Taiwan

The significance of latex paint as a source of indoor volatile or-ganic compounds is underscored by the large volume producedfor interior use. This review focuses on one important compo-nent of latex paint, 2,2,4-trimethyl-1,3-pentanediol monoisobu-tyrate (TMPD-MIB). Past research is described, with an emphasis onmeasurements of TMPD-MIB emissions, experimental recoveries ofTMPD-MIB, and factors that affect the emissions process. Publishedmodels that attempt to describe TMPD-MIB emissions following latexpaint applications are summarized. Finally, a critical assessmentof the state of knowledge related to TMPD-MIB emissions from latexpaints is presented, along with a summary of continuing researchneeds.

KEY WORDS: Texanol©R , fate, architectural coatings, gypsumboard, models

INTRODUCTION

Architectural coatings applied to building materials (substrates) are a sourceof emissions of volatile and semi-volatile organic compounds ((S)VOCs) toboth indoor and outdoor atmospheres. Concerns related to such emissionsare two-fold. First, many of the constituents of architectural coatings are

Address correspondence to Richard L. Corsi, Department of Civil, Architectural and En-vironmental Engineering, The University of Texas, 1 University Station (C1786), Austin, TX78712, USA; Tel.: (512) 475-8617; Fax: (512) 471-1720; E-mail: [email protected]

1052

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

Emissions of TMPD-MIB from Latex Paint 1053

reactive and can contribute to the production of photochemical smog inurban air sheds (Klimont et al., 2002; Lawrimore & Aneja, 1997; Waddenet al., 1994). As such, the inclusion of speciated (S)VOC emissions fromarchitectural coatings in urban emissions inventories is desirable. Second,architectural coatings applied indoors can lead to elevated (S)VOC concen-trations in buildings, relative to the outdoor environment, over short periodsof time (days), and may persist at much lower levels for months to years(Clausen et al., 1991; Hodgson, 1999; Hodgson et al., 2000; Lin & Corsi, 2007;USEPA, 2002). Sparks et al. (1999a) noted the large surface areas coveredby indoor paints, primarily water-based paints, and the potential impacts onindoor air quality.

The significance of paint products as sources of indoor (S)VOCs is un-derscored by the large volume produced for interior use. Approximately 795million gallons of architectural coatings were produced in the United Statesin 2005 (United States Department of Commerce, 2006). Of this, 65% (519million gallons) was classified as being for interior use. Of the interior-typecoatings, 89% (460 million gallons) was classified as interior water-type, asopposed to solvent-type coatings. Water-type paints, often referred to as la-tex paints and the focus of this review, comprised 84% (385 million gallons)of interior water-type coatings, with an approximate equal split between flatwater-thinned paints/tinting bases and semi-gloss, eggshell, satin, and otherwater-thinned paints/tinting bases. However, despite the large amounts ofarchitectural coatings now being produced and used around the world, aswell as the potential impacts of emissions from architectural coatings on out-door and indoor air quality, there is still a great deal of uncertainty regardingboth the impacts of such emissions and the nature of emissions processes.

This review focuses on one important component of latex paints, 2,2,4-trimethyl-1,3-pentanediol monoisobutyrate, often referred to as Texanol©R es-ter alcohol and hereafter denoted TMPD-MIB. There are several compellingreasons to review the existing knowledge base related to TMPD-MIB, thefirst of which is the shear mass production rate and fact that the public gen-erally has close contact with the chemical given its use in paint products. Theglobal production rate of TMPD-MIB is estimated at 98,000 metric tons/year(Rector, 2005). Further, TMPD-MIB has a low-odor threshold concentrationand is thus often detected for hours to days after a painting event has oc-curred. It has an airway irritation threshold concentration that is reasonablyachieved when painting indoors.

Peer-reviewed publications related to the toxicological effects of TMPD-MIB in humans is sparse to non-existent, although tests for mutagenic-ity using the Ames assay and several strains, as well as in vivo mouse,micronucleus assays proved negative (Nielsen et al., 1997). One recentstudy indicated a positive association between TMPD-MIB concentrationsin school classrooms and nocturnal breathlessness in the home (Kim et al.,2007). However, the home environment was not characterized in the study.

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

1054 R. L. Corsi and C.-C. Lin

TMPD-MIB has come under some scrutiny as a precursor to urban ozoneformation. Recent studies indicate a maximum incremental reactivity (MIR)of TMPD-MIB equal to 0.88 g ozone/g TMPD-MIB reacted (Carter & Malkina,2005). As such, TMPD-MIB may contribute to the formation of ozone in ur-ban air sheds. Finally, there is little in the published literature regarding thefate of other oxygenated chemicals that are used in architectural coatings,particularly ethylene and propylene glycol, and 2-(2-butoxyethoxy) ethanol(BEE). Completion of a review of TMPD-MIB may shed light on the fate ofother oxygenated components used in architectural coatings and cleanersand facilitate future reviews or experimental studies of those compounds.

In this review, background information is provided to establish a ba-sis for attempting to understand TMPD-MID emissions from latex paintsand potential improvements in the existing knowledge-base related to suchemissions. Past experimental research is described, with an emphasis onmeasurements of TMPD-MIB emissions following latex paint applications,experimental recoveries of TMPD-MIB, and factors that affect the emissionsprocess. Published models that attempt to describe TMPD-MIB emissions fol-lowing latex paint applications are summarized. Finally, a critical assessmentof the state of knowledge related to TMPD-MIB emissions from latex paintsis presented, along with a summary of continuing research needs.

BACKGROUND

Latex Paints

Latex paints are water-thinned paints that contain significantly lower amountsof VOCs than oil-based paints, and tend to dry much more rapidly. Inter-estingly, water-based paints have been used since pre-industrial times. Forexample, from 1500 to 1000 B.C., Egyptians used water-based paints com-prised of indigo and mud, with gum Arabic as a binder (Norback et al., 1995).Today, most paints, including latex paints, are comprised of four major com-ponents: pigments, binders, solvents, and additives. Pigments consist of fineparticles that are used to provide color and opacity or covering power. Theparticles are generally inorganic in nature (e.g., titanium dioxide). Bindersassist with binding of pigments and additives together and provide adhesionto the substrate being coated. Higher performance latex paints often includeacrylic resins resulting from the polymerization of derivatives of acrylic acidsas a binder. However, other binders such as polyvinyl acetate are also em-ployed in latex paints. Solvents aid in maintaining unused paint in a liquidform prior to application and allow pigment and binder solids to behaveas a fluid during application. Solvents are intended to evaporate completelyfollowing application, with evaporation rates largely dependent on the na-ture of the solvent mixture and indoor environmental conditions (e.g., airspeed, humidity, temperature). Water and glycol solvents, principally ethy-lene glycol and propylene glycol, are used in latex paints (i.e., as opposed to

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

Emissions of TMPD-MIB from Latex Paint 1055

aliphatic and aromatic hydrocarbons used in oil-based paints). The glycolsadded to latex paint also provide for freeze protection. Most paints also con-tain one or more additives that serve special purposes. These additives caninclude biocides, coalescing agents, defoaming agents, thickeners, and sta-bilizers, among others. A key additive in most latex paints, and the focus ofthis paper, is 2,2,4-trimethyl-1,3-pentanediol monoisobutyrate (TMPD-MIB).TMPD-MIB is added as a coalescing aid. It is adsorbed and/or absorbed by,and helps to soften, polymeric binder particles, a property that facilitatescomplete fusion when paint dries. It also enhances scrub resistance, colordevelopment, and packaging stability.

Volatile Components of Latex Paint

The percentage by weight of the non-volatile components (resin + pigment)of latex paint has been reported to be in the range of 50–60%, and a per-centage by weight water content of 40–50% appears common (e.g., Changet al., 1997; Fortmann et al., 1993; Sparks et al., 1999a). Water dominates theoverall percentage by weight volatile content. The total volatile organic com-pound (TVOC) content typically ranges between 2–5% by weight (USEPA,2002). Hereafter, TVOC will refer to the sum total of compounds classifiedas both volatile and semi-volatile organic compounds.

The actual composition of TVOC can vary significantly between latexpaints (e.g., semi-gloss latex paints are typically comprised of greater TMPD-MIB weight fractions than are flat latex paints). Censullo et al. (1996) reportedTVOC composition analysis for 34 latex paints. They observed a wide rangeof (S)VOCs and differing (S)VOC ratios over all paint samples. However,four (S)VOCs clearly dominated in contribution to TVOC: diethylene gly-col butyl ether (also commonly reported as 2-(2-butoxyethoxy)ethanol, orBEE), ethylene glycol (EG), propylene glycol (PG), and TMPD-MIB. Basedon their analysis, Censullo et al. (1996) proposed the following compos-ite group species profile for water-based paints as percentage by weight ofTVOC: TMPD-MIB (35%), PG (33%), EG (17%), BEE (6%), other (residual).However, this was intended as a gross species profile that could be used foroutdoor emissions inventories. It does not account for the significant vari-ability between waterborne paints. For example, the percentage by weightTMPD-MIB contribution to TVOC varied from 0% to greater than 70%.

Several other researchers have reported TMPD-MIB of between 20 and33% of TVOC on a percentage by weight basis for latex paints intended forinterior applications (Chang et al., 1997; Fortmann et al., 1993; Sparks et al.,1999a, 1999b; Wilkes et al., 1996). This range is slightly less than, but closeto, the central tendency reported by Censullo et al. (1996). In each of theaforementioned studies, the dominant (S)VOCs were consistent with thosereported by Censullo et al. (1996). Ethylene glycol and TMPD-MIB, in thatorder, dominated TVOC composition, with lesser amounts of BEE and PG

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

1056 R. L. Corsi and C.-C. Lin

in each case. The reported contribution of EG was much greater, and thereported contribution of PG much lower, than that reported by Censullo etal. (1996).

Assuming a TMPD-MIB content of 20–33% by weight of TVOC and aTVOC content of 2–5% by weight yields an estimated TMPD-MIB contentof approximately 0.4–1.7% by weight of paint. Sheldon and Naugle (1994)analyzed six interior latex paints and reported a TMPD-MIB range of 3–10 g/kg (0.3–1% by weight). They also reported EG (0.6–4.8%) and BEE(0.15–1.3%). Hodgson (1999) reported the compositions of a latex primersealer, flat latex paint, and semi-gloss latex paint used in large chamberexperiments. The percentage by weight compositions of TMPD-MIB for thesethree products was 1.5%, 0.9%, and 0.7%, respectively. The percentage byweight compositions of EG were 3.6%, 2.0%, and <0.7%, respectively. Thepercentage by weight compositions of PG were <0.8%, <0.8%, and 3.1%,respectively. Fortmann et al. (1993) reported the mean composition of threereplicates of a single latex paint as TMPD-MIB (0.75% by weight), EG (1.3%),and BEE (0.5%). Chang et al. (1997) reported the mean composition of fourreplicates of a single latex paint as TMPD-MIB (1.35% by weight), EG (2.4%),PG (0.24%), and BEE (0.5%). This specific composition/paint served as thebasis for experiments/model development by the USEPA, which will bedescribed later in this review (Sparks et al., 1999a, 1999b).

Properties of TMPD-MIB

Several properties of TMPD-MIB are listed in Table 1. For purposes oflater discussion, the properties of several other components of latex paint

TABLE 1. Properties of major volatile components of latex paint.

CompoundMolecularformula

CAS#

MW[g/mol]

Aqueoussolubility[mg/L]@ 20◦C

Pvp

[mmHg]

@ 20◦CTb

[◦C]Tf

[◦C]log10

(Kow)

BEE C8H18O3 112-34-5 162.2 Miscible 0.02 231 −68 0.15 to 0.40EG C2H6O2 107-21-1 62.1 Miscible∗ 0.05 198 −12.7 −1.93PG C3H8O2 57-55-6 76.1 Miscible∗ 0.2 188.2 −59 −1.40 to −0.30TMPD- C12H24O3 25265-77-4 216.4 858 0.01 254 −50 3.47

MIB (18–22◦C)Water H2O 7732-18-5 18 — 17.54 100 0 —

For each chemical, the relevant temperature was not reported for Kow. Values for water are widelytabulated and published extensively.All values except for TMPD-MIB and water are from Verschueren (1996) and ∗Agency for Toxic Substancesand Disease Registry [ATSDR] (1997). Values for TMPD-MIB taken from Eastman Chemical Company(2003).Abbreviations: MW = molecular weight, Pvp = vapor pressure, Tb = boiling point, Tf = freezing/meltingpoint, Kow = octanol/water partition coefficient.

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

Emissions of TMPD-MIB from Latex Paint 1057

are also listed in Table 1. TMPD-MIB is a mixture of two isomers (2,2,4-trimethyl-1,3-pentanediol-x-isobutyrate, where x = 1 or x = 3). In one of theisomers (x = 1), the ester bond involves the hydroxy group of the C1 atomin the diol moiety, and in the second isomer (x = 3), it involves the hydroxygroup of the C3 atom. The mixture is a colorless liquid with a mild odor.The odor threshold has been reported to be 66 ppb (Ziemer et al., 2000),and the airway irritation threshold has been noted to be 1,000 µg/m3 (≈112ppb at 25◦C) (Knudsen et al., 1999, and references provided therein).

Three important physico-chemical properties of TMPD-MIB are its rel-atively low vapor pressure, low solubility, and high octanol-water partitioncoefficient. The former indicates a relatively slow rate of evaporation. Thelatter indicates a strong tendency for accumulation into or onto organic mate-rial (e.g., organic resins in latex paint) and removal from the aqueous phase.TMPD-MIB is often referred to as a semi-volatile organic compound (SVOC)due to its relatively high boiling point.

Emissions of TMPD-MIB from Latex Paints: Conceptual Development

It is worthwhile to consider the major processes that may affect TMPD-MIBemissions, even in qualitative terms, as a prelude to the interpretation ofpublished research results described later in this paper.

The release of TMPD-MIB from latex paint to indoor air is a dynamicprocess that is influenced by the continuously changing nature of the paintfollowing application to a material (i.e., from initial wet paint compositionto ultimate dry paint film). However, the emissions process is generallydescribed by two major stages. The first is a relatively short stage, duringwhich the paint is treated as a wet film. The second is a prolonged stagein which the paint is treated as a dry film. These two stages are depicted inFigure 1. In the wet film, TMPD-MIB can exist in three major components:dissolved in the aqueous phase, as a separate pure phase, and adsorbedto (or absorbed by) resin particles. It may also volatilize to room air, orpartition into the gas phase with subsequent diffusion into pores of thesubstrate to which the paint is applied. The presence of co-solutes such asethylene glycol (EG), propylene glycol (PG), and 2-(2-butoxyethoxy)ethanol(BEE) may have some influence on the solubility of TMPD-MIB in the liquidportion of the paint emulsion. However, the solubility of TMPD-MIB is likelyto remain very low. As described previously, the percentage by weight ofTMPD-MIB in latex paint is generally in the range of 0.4–1.7% (4–17 g/kg).For a latex paint in which water comprises 50% of total weight, this rangewould translate to 8,000 to 34,000 mg/L if all of the TMPD-MIB could bedissolved in water, values that exceed the aqueous solubility limit of TMPD-MIB by a factor of approximately 9–40. Thus, it is expected that a very smallfraction of the TMPD-MIB mass in paint is dissolved in the liquid phase.

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

1058 R. L. Corsi and C.-C. Lin

FIGURE 1. Schematic depicting latex paint applied to a building material (substrate), includ-ing emissions from (a) a wet film and (b) ultimate dry film absent of water. The lines internalto the substrate depict VOC fate mechanisms internal to the substrate, including moleculardiffusion, adsorption, and reaction mechanisms.

As reflected by the very high octanol-water partition coefficient of TMPD-MIB, and by its intent as a coalescing additive that facilitates the softeningof resin particles, it is reasonable that a majority of its mass exists adsorbedto, or absorbed within, the large mass of organic resin present in the paint.Nevertheless, the fraction dissolved in the aqueous phase and any separatepure phase will play some role in the emissions process during the drying(wet film) stage. The mechanistic behavior of corresponding emissions iscomplicated by the fact that water evaporates faster than the other solventsor TMPD-MIB, thus leading to changes in co-solvent effects and relativepartitioning of TMPD-MIB between the liquid and particle phases of theemulsion.

The dry film stage (see Figure 1b) will exist following complete evapo-ration of water. The dry film is comprised of resin-pigment, residual organicsolvents, and TMPD-MIB. The process by which TMPD-MIB emissions occurfrom the dry film are not well understood, but have been reported to persistlong after paint drying (Clausen et al., 1991; Hodgson et al., 2000; Lin &Corsi, 2007; USEPA, 2002), and a majority of emissions have been noted tooccur from the dry film (USEPA, 2002). It is likely that molecular diffusionthrough the dried paint film, and possibly the substrate (Lin & Corsi, 2007),plays a major role in prolonged emissions from paint after it reaches thedry film stage. Clausen et al. (1991) hypothesized that either diffusion orevaporation can dominate emissions after the film forms, depending on thespecific volatile composition of the paint and interactions between (S)VOCs.

The TMPD-MIB that diffuses into the substrate volume is likely to inter-act significantly with the substrate (e.g., gypsum board or concrete). There is

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

Emissions of TMPD-MIB from Latex Paint 1059

only one published study for which TMPD-MIB storage in a material (gypsumboard) was evaluated (Lin & Corsi, 2007). However, there are no publisheddata to allow for estimates of TMPD-MIB storage by adsorption to othermaterials, such as engineered wood products or concrete. As such, the de-velopment of a diffusion/adsorption model to describe the fate of TMPD-MIBwithin these substrates has not been effectively formulated. Also, there aresparse data to allow for an assessment of irreversible chemi-sorption withinthe substrate, or reactions within the dry paint film.

SUMMARY OF PAST RESEARCH

Empirical observations of TMPD-MIB emissions from latex paint have beenreported in the published literature. Much of this is related to the use of smallcontinuous flow laboratory chambers constructed of stainless steel or glass(Chang et al., 1997; Clausen, 1993; Clausen et al., 1991; De Bortoli et al., 1999;Fang et al., 1999; Fortmann et al., 1993; Hodgson, 1999; Hodgson & Shimer,1999; Knudsen et al., 1999; Lin & Corsi, 2007; Roache et al., 1996; Sparks etal., 1999a; Van der wal et al., 1997) or the Field and Laboratory EmissionCell (FLEC) (De Bortoli et al., 1999; Roache et al., 1996; Wolkoff, 1998)to quantify area-specific emission rates. The relevance of many laboratoryexperiments to real-world conditions has been limited due to the extensiveuse of impermeable (non-porous) substrate materials such as stainless steel,aluminum, glass, and plexi-glass.

Hodgson (1999) completed large chamber experiments intended to sim-ulate conditions in actual residential room environments. A limited numberof field studies/experiments have also been completed (Hodgson et al., 2000;Sparks et al., 1999a). These studies have shed some light on TMPD-MIB con-centrations in actual indoor environments that are influenced by sorptivesinks (i.e., other materials to which TMPD-MIB can adsorb to and desorbfrom).

Every study reviewed for this paper, whether conducted in the labora-tory or field, involved the collection of TMPD-MIB on a solid adsorbent withsubsequent thermal or chemical extraction and analysis via either GC/FID orGC/MS. Tenax©R -TA has been by far the most widely used sorbent, with sub-sequent extraction of TMPD-MIB by thermal desorption (Chang et al., 1998;Clausen, 1993; Clausen et al., 1991; Knudsen et al., 1999; Hodgson & Shimer,1999; Hodgson et al., 2000; Lin and Corsi, 2007; Roache et al., 1996; Sparkset al., 1999a; Yu & Crump, 1999). Van der wal et al. (1997) employed char-coal tubes with solvent extraction using carbon disulfide (CS2). However,Norback et al. (1995) reported poor recovery of TMPD-MIB from standardcharcoal tubes when extracted by CS2. Fortmann et al. (1993) used XADresin extracted with methylene chloride, but noted the need for more per-formance data related to different sorbents. De Bortoli et al. (1999) reported

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

1060 R. L. Corsi and C.-C. Lin

very good inter-laboratory results for TMPD-MIB emissions associated withan 18 laboratory/10 country European study involving three materials, oneof which was a latex paint. However, they concluded that analytical errorswere by far the greatest cause for variance in results between laboratories.

Past research has provided some insight into the order of magnitudeof TMPD-MIB emissions following latex paint applications to both idealizedand typical building materials (De Bortoli et al., 1999; Fang et al., 1999;Hodgson, 1999; Lin and Corsi, 2007; Roache et al., 1996; among others).The mass recovery of TMPD-MIB in air following paint application has beenreported by several research teams (Chang et al., 1997, 1998; Guo et al., 1996;Lin and Corsi, 2007; among others). There has also been some research onthe effects of the method of application and subsequent film thickness onTMPD-MIB emissions (Clausen, 1993; Clausen et al., 1991; Fortmann et al.,1993; Yu & Crump, 1999). Several research teams have reported on theeffects of environmental conditions on TMPD-MIB emissions, with variationsin air temperature, relative humidity, and air speed above the wet and/ordry paint film (Fang et al., 1999; Hodgson & Shimer, 1999; Knudsen et al.,1999; Roache et al., 1996; Van der wal et al., 1997; Wolkoff, 1998). Relevantfindings from these past studies are summarized in the following sections.

TMPD-MIB Concentrations in Indoor Environments

There is limited information regarding TMPD-MIB concentrations, in eitherindoor or outdoor air, in the published literature. Girman et al. (1999) re-ported VOC concentrations in 56 public and private office buildings as partof the USEPA’s Building Assessment Survey and Evaluation study (BASE).TMPD-MIB was among the most frequently detected compounds (81–100%frequency of quantifiable concentrations over all samples). The range ofTMPD-MIB concentrations was 0.5 to 28 µg/m3 (∼0.1–3 ppb). Hodgsonet al. (2000) completed a field study involving 11 homes (four manufac-tured homes and seven site-built homes). The four manufactured homeswere sampled between 2 and 9.5 months after installation. The concentra-tions of TMPD-MIB in these homes were as follows: range = 1.4–6.7 ppband geometric mean (GM) = 2.4 ppb, with geometric standard deviation(GSD) = 1.5 ppb. The seven site-built homes were sampled 1–2 months af-ter completion. The concentrations of TMPD-MIB in these homes were asfollows: range = 3.1–25.1 ppb and GM = 9.1 ppb, with GSD = 2.2 ppb.

Sparks et al. (1999b) conducted sampling in a USEPA test house afterpainting the walls of a bedroom with latex paint. All of the walls were firstcovered with new gypsum board. Latex paint was then applied with a rollerto an area-average mass coverage of 155 g/m2. The average air exchange rateof the test house was 0.37/hr, with a temperature of 22.2◦C and a relativehumidity of 70%. The approximate (read from log axis) concentration of

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

Emissions of TMPD-MIB from Latex Paint 1061

TMPD-MIB at 200 and 1400 hours (after paint application) in the paintedbedroom was 10 ppb and 3 ppb, respectively. Similar concentrations wereobserved in a hallway near the bedroom at the same times. The value of 3ppb nearly two months after paint application was generally consistent withthe ranges reported by Hodgson et al. (2000) for residential buildings.

Shields and Weschler (1992) completed a three-year study of VOCs ina telephone switching station in Wisconsin, USA. TMPD-MIB was observedin room air during and following two renovation events in the building. Themeasured concentrations ranged from highs of approximately 1 to 2 ppbduring renovation to between 0.6 and 1.1 ppb three months later. Theserelatively low values were likely due to high ventilation rates in the build-ing during and following renovation events. During the first and secondrenovation events, supply air was comprised of 37% and 68% outside air,respectively.

Norback et al. (1995) completed an exposure analysis of 20 housepainters in Sweden. Exposure concentrations were measured during the ap-plication of waterborne paints in several types of building environments.However, the TMPD-MIB contents of the various paints were not reported.The maximum exposure concentration to TMPD-MIB was observed to be1,680 µg/m3(≈190 ppb at 20◦C and 1 atmosphere). The arithmetic mean ex-posure concentration was 164 µg/m3 (≈19 ppb at 20◦C and 1 atmosphere)and the geometric mean exposure concentration was 18 µg/m3 (≈2 ppbat 20◦C and 1 atmosphere) (GSD = 7.0 µg/m3). Thus, while some painterswere exposed to relatively high concentrations of TMPD-MIB, the geomet-ric mean exposure concentration was on the same order of magnitude asbackground concentrations reported by Girman et al. (1999) for commer-cial office buildings and Hodgson et al. (2000) and Sparks et al. (1999b) forresidential homes several months after painting.

Kim et al. (2007) measured TMPD-MIB concentrations in 23 classroomsof eight schools in Sweden. The mean concentration was 0.89 µg/m3, witha range of 0.07 to 4.41 µg/m3. The highest concentrations were measured innewer schools.

Finally, Hodgson and Shimer (1999) completed large chamber experi-ments with material assemblies and conditions designed to simulate an in-door residential room environment. The sampling periods were closer to thetime of application than those studies described above. A latex primer sealerwas applied to gypsum board and plywood. Flat and semi-gloss latex paintswere then applied to gypsum board and plywood, respectively. The basecondition had an air exchange rate of 2/hour for several hours, followedby an extended period of air exchange at 0.5/hr. Chamber air exchangerates and air mixing intensity were varied, but had little effect on measuredconcentrations of TMPD-MIB. Over three experiments, the TMPD-MIB con-centration ranged from 480 to 630 ppb during 0 to 48 hours, and from 138 to179 ppb at 240 hours, levels much greater than those reported by Norback

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

1062 R. L. Corsi and C.-C. Lin

et al. (1995) for painter exposure concentrations during water-based paintapplications.

Emissions of TMPD-MIB from Latex Paint

Published measurements of TMPD-MIB emissions following latex paint ap-plications are sparse and generally limited by a lack of analysis of initial paintcomposition, as well as numerous reports of paint applications to imperme-able substrates such as stainless steel and glass. One of the most comprehen-sive and well-documented studies to date was completed by Hodgson (1999).Data extracted from the original source report are summarized in Table 2.The reported area-specific and mass-specific emission factors are based onflow-through small chamber experiments and were calculated based on anassumption of pseudo-steady-state conditions at the time of sampling (48 and96 hours after product application). The highest emission factors were for alatex primer sealer and flat latex paint applied to gypsum board. Emissionfactors for paints applied to gypsum board decreased significantly from hour49 to 96. This was not the case for semi-gloss paint applied to plywood.

Emission factors gleaned from six other published studies are reportedin Table 3. Four of these involved applications of latex paints to gypsumboard, but are not comparable in time to the results presented in Table 2.

TABLE 2. TMPD-MIB emission factors extracted from Hodgson (1999) for small chamberexperiments.

Coverage Emission factor Emission factorof wet (mg/m2-hr) (mg/kg-hr)paint

Product Substrate (g/m2) 48 hours 96 hours 48 hours 96 hours

LPS1 Gypsum board 79 0.94 0.55 11.9 6.94LPS2 Gypsum board 110 2.64 1.97 24.0 18.0FLP1a Gypsum board 99 2.02 0.26 20.8 2.59FLP1b Gypsum board 104 2.32 0.30 22.4 2.93FLP2 Gypsum board 111 1.43 0.28 12.9 2.48FLP3 Gypsum board 163 2.86 1.42 23.8 8.73SGLP1 Plywood 164 0.34 0.26 2.07 1.56SGLP2 Plywood 129 0.88 0.82 6.88 6.33SGLP3a Plywood 169 0.93 0.89 5.55 5.30SGLP3b Plywood 160 0.72 0.54 4.39 3.33LPS2 + FLP3 RH = 30% Gypsum board 187 5.61 3.83 30.7 20.9LPS2 + FLP3 RH = 50% Gypsum board 156 6.26 3.20 40.0 20.5LPS2 + FLP3 RH = 70% Gypsum board 163 5.32 2.68 32.7 16.5LPS2 + SGLP3 Plywood 176 1.32 0.76 7.49 4.33

SGLP was applied with a brush to smooth plywood. FLP was applied with a roller to gypsum board.Chambers are constructed of 316 stainless steel, volume = 10.5 L, air exchange rate = 5.7 ± 0.3/hr,temperature = 23 ± 1◦C, RH (standard) = 50 ± 5%, air velocity near surface of substrate ≈0.25 m/s,loading ratio = 1.86.Abbreviations: LPS = latex primer sealer; FLP = flat latex paint; SGLP = semi-gloss latex paint.

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

Emissions of TMPD-MIB from Latex Paint 1063

TABLE 3. TMPD-MIB emission factors extracted from various sources.

Emission factor TimeMaterial mg/m2-hr (units noted) Note Reference

Gypsum board 0.002 11 months Estimated based onread off logconcentration plot

Chang et al.(1997)

Gypsum board 0.017 1,500 hours Roache et al.(1996)

Gypsum board 0.0041–0.0046 1,500 hours FLEC Roache et al.(1996)

Gypsum board 0.027–0.216 3 weeks Varied RH & T Fang et al. (1999)Several

impermeablesubstrates

4.8 48 hours 18 lab study: CV =21–25%, smallchambers andFLEC used

De Bortoli et al.(1999)

Glass 29 3 hours Yu and Crump(1999)

3.3 27 hours1.4 76 hours

Nevertheless, it appears that significant reductions in area-specific emissionrates for TMPD-MIB occur over time. The last two entries in Table 3 involveapplications of latex paints to impermeable substrates such as stainless steel,polyester sheets, and glass. The time scales are similar to those reported byHodgson (1999) for gypsum board. Interestingly, the area-specific emissionrates for the impermeable materials for the first 27 to 48 hours are reasonablysimilar to those reported by Hodgson (1999) for gypsum board, although adirect comparison of results is difficult without knowing the actual com-position of the paints used in each study. For an impermeable substrate(glass), Fortmann et al. (1993) observed similar concentrations in a small testchamber at 4 and 24 hours after paint application.

Long-term emissions data for TMPD-MIB are sparse. Lin and Corsi (2007)applied latex paint to gypsum board and measured concentrations in a smallflow-through chamber over a 15-month period. The concentration of TMPD-MIB decreased significantly within the first 100 hours after paint application.Thereafter, low but measurable and relatively constant concentrations wereobserved for up to 15 months.

Chang et al. (1997) applied latex paint to gypsum board and measuredconcentrations in a small flow-through chamber over an 11-month period.The concentration of TMPD-MIB was greater than that of ethylene glycol forthe first 100 hours and was lower thereafter. Significantly, TMPD-MIB, EG,and PG were all detected in chamber air 11 months after paint application.

Clausen et al. (1991) applied five different water-based paints to tin-plated stainless steel sheets. TMPD-MIB was the only compound detected inoff-gas after one year.

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

1064 R. L. Corsi and C.-C. Lin

Sparks et al. (1999b) applied latex paint to gypsum board in a USEPAtest house. TMPD-MIB concentrations in house air decreased significantlyover the first 400 hours. However, they were relatively constant from 400 to4000 hours after application, suggesting either relatively constant prolongedemissions from the dry paint and/or constant re-emission rates from interiorsinks that had adsorbed TMPD-MIB over the field study.

Experimental Recoveries of TMPD-MIB

The recovery of a compound following application of paint to a substrate istypically defined as the cumulative mass emitted to air divided by the mass ofthe compound initially applied to the substrate. The recovery of total volatileorganic compound (TVOC) mass two weeks after latex paint is applied tostainless steel has been observed to be nearly 100% (Sparks et al., 1999a, andreferences provided therein). Similarly, recovery of TMPD-MIB two weeksafter latex paint is applied to stainless steel has been observed to be on theorder of 90% (Chang et al., 1997; Guo et al., 1996).

Only a few TMPD-MIB recoveries have been reported for latex paintapplications to gypsum board; no published reports of TMPD-MIB recoveryfollowing latex paint applications to wood, concrete, or stucco were identi-fied during this review. A summary of published results for painted gypsumboard is presented in Table 4. Note that the recovery period never exceededtwo weeks in these studies except for Lin and Corsi (2007) and, as such,long-term recoveries of TMPD-MIB from actual building materials are stilldeficient. Over the three two-week studies, the range of TMPD-MIB recov-eries from gypsum board is relatively large (27 to 60%), with differencesbeing difficult to ascertain based on available data. All three two-week pe-riod recoveries were based on latex paint applied to gypsum board placedin small stainless-steel flow-through chambers with similar air exchange rate,relative humidity, and temperature. The compositions of the paints used byChang et al. (1997) and Guo et al. (1996) were similar in terms of TVOC andpercentage by weight contribution of TMPD-MIB and other (S)VOCs. Thecomposition of paint used by Roache et al. (1996) was not described in thesource paper.

TABLE 4. Recovery of TMPD-MIB mass following the application of latex paint to gypsumboard.

Sampling period(weeks) Recovery (%) Reference

2 29 Chang et al. (1997)2 27–33 Roache et al. (1996)2 60 Guo et al. (1996)64 50–90 Lin and Corsi (2007)

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

Emissions of TMPD-MIB from Latex Paint 1065

Lin and Corsi (2007) observed TMPD-MIB emission rates for periodsas long as 15 months. The airborne recoveries of TMPD-MIB were a strongfunction of the type of paint and substrate. Recoveries in air of approximately50% (low-pigment volume paint) to 90% (high-pigment volume paint) wereobserved after 15 months for regular applications to gypsum board. Lin andCorsi (2007) were the first to complete solvent extractions of the paintedsubstrate and to perform a mass closure analysis of TMPD-MIB. Resultsfor 10 gypsum board specimens yielded a 96 ± 6% mass closure (i.e., onaverage, nearly all of the TMPD-MIB not emitted to air was typically observedin the substrate extracts). A fairly high percentage (≈40%) of the TMPD-MIB applied to gypsum board in low pigment volume latex paint (tendingtoward semi-gloss paint) was sorbed within the gypsum board matrix, even15 months after paint applications. Low-level emissions to air after threemonths following paint applications were dominated by releases from thedry paint film, and not in the form of TMPD-MIB diffusing from the interiorof the gypsum board.

Based on a review of existing literature, there is a clear need for morelong-term TMPD-MIB mass recovery data following applications to productsother the gypsum board, including wood, concrete, plaster, and stucco.

Effects of Substrate on TMPD-MIB Emissions from Latex Paint

Early studies related to (S)VOC emissions from latex paint were based onimpermeable (non-porous) materials such as stainless steel, glass, and alu-minum (Clausen, 1993; Clausen et al., 1991; Fortmann et al., 1993). Evenmore recently, several researchers have continued to use substrates such asglass (Yu and Crump, 1999), aluminum (Van der wal et al., 1997), polyestersheets, stainless steel, and aluminum (De Bortoli et al., 1999). In fairness,the intent of these studies was not necessarily to provide realistic emissionsdata for indoor paint applications; for example, the study by De Bortoliet al. (1999) was intended as an inter-laboratory comparison of methodsused to characterize emissions from building materials.

More recently, there has been greater emphasis on the use of actualindoor substrates to study emissions from latex paint. Most of the publishedaccounts of latex paints applied to actual building materials have focusedon gypsum board as the substrate (Chang et al., 1997; Fang et al., 1999;Guo et al., 1996; Hodgson, 1999; Knudsen et al., 1999; Lin & Corsi, 2007;Wolkoff, 1998), and not all of these have provided sufficient information todetermine emission rates. For this review, only one set of results for latexpaint applications to a wood product (plywood) was found (Hodgson, 1999).

The substrate to which paint is applied can have a significant influenceon component emissions. Gehrig et al. (1993) may have been the first to re-port differences in (S)VOC emissions from a single paint applied to different

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

1066 R. L. Corsi and C.-C. Lin

substrates, including glass, gypsum board, and gypsum board covered withwood chip wallpaper. Substrate effects were relatively small for the non-polarn-alkanes and aromatics. Although TMPD-MIB was not included in the paintmixture, the substrate did have a significant influence on two polar com-pounds, BEE and 2-(2-butoxyethoxy) ethanol acetate. For each compound,a significant reduction in emissions was observed following applications toeach of the two gypsum board substrates relative to emissions from glass.Krebs et al. (1995) also emphasized that substrates can have a significantinfluence on emissions from paints.

Comparisons have been made to determine differences in (S)VOC emis-sions from the same paint applied to gypsum board and stainless steel (Changet al., 1997, Guo et al., 1996; Lin & Corsi, 2007; Sparks et al., 1999a). Signifi-cant differences in the concentration-time profile for either TVOC or individ-ual components of TVOC, including TMPD-MIB, between the two substrateshave been observed. In general, the peak concentration in chamber air oc-curs sooner and is smaller when paint is applied to gypsum board; also,as noted previously, mass recovery of TVOC or individual components arelower over the experimental period for such applications.

The fact that emissions are significantly influenced by choice of sub-strate in the first few hours of small chamber experiments indicates that thesubstrate even plays a significant role during the wet-film phase of the emis-sions process. Because this effect appears to depend on the polarity of paintcomponents, the emissions retardation cannot simply be due to shieldingof mass from evaporative processes by liquid paint “soak” into the pores ofa substrate such as gypsum board. Instead, emissions may be reduced bysorptive interactions between polar compounds and gypsum board whilein either the liquid phase (liquid-solid interactions) or following liquid-gaspartitioning and gas-phase diffusion/adsorption processes within the poresof the substrate.

Importantly, it appears that the porosity and/or polarity of a substrate isan important factor in determining oxygenated (S)VOC emissions followingpaint application. Applications of the same latex paint to different non-porous materials (polyester sheets, glass, aluminum) had little impact onarea-specific emission rates during a large inter-laboratory comparison (DeBortoli et al., 1999).

The nature of gypsum board also seems to have little influence on(S)VOC emissions following paint applications. Sparks et al. (1999a) de-scribed a study in which the same latex paint was applied to new gypsumboard, gypsum board that had been previously painted, and eight-year-old,painted gypsum board. Emissions of TVOC following fresh paint applicationswere similar in each case.

Silva et al. (2003) appear to be the first to report on emissions of TMPD-MIB and other compounds following the application of a latex paint to con-crete. In direct comparison with application of the same paint to polyester

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

Emissions of TMPD-MIB from Latex Paint 1067

sheets, the emission rates were significantly lower initially (wet-film phase)for concrete, but greater for concrete during the dry-film phase; experimentswere 72 hours in length. Retention was so significant on concrete for BEE,diethyl phthalate, and TMPD-MIB that the authors speculated on the pos-sibility of either irreversible or very slow reversible sorption processes forthese compounds. These results are significant and warrant further researchgiven extensive applications of paint to concrete in some parts of the world.For example, in China, a majority of architectural paint products, includinglatex paints, is applied to concrete (Avery, 2004).

It is clear that the choice of substrate has a significant influence onemissions from latex paint, and that the emissions process is retarded forTMPD-MIB and other polar components when latex paint is applied to gyp-sum board. While more research is needed to fully understand the nature ofthe emissions process following paint applications to gypsum board, woodproducts and concrete are also deserving of additional attention. Only onestudy was identified for each of these substrates.

Effects of Latex Paint Application Method on TMPD-MIB Emissions

The use of rollers or brushes is preferred for application of paints to testsubstrates for emissions testing (USEPA, 2002). Most published studies haveinvolved the application of latex paint to gypsum board with rollers (Changet al., 1997; Roache et al., 1996; Sparks et al., 1999a; Knudsen et al., 1999;Lin & Corsi, 2007). Fortmann et al. (1993) and Yu and Crump (1999) usedbrushes to apply latex paint to glass.

The most important factor associated with paint application appears tobe the thickness of the paint film. Clausen et al. (1991) found that long-termemissions are reduced by reducing the paint film thickness on tin-platedsteel, and that normalization of the emissions by film thickness is criticalfor avoiding incorrect conclusions when comparing emissions from paint.For TMPD-MIB, peak chamber concentrations increased with increasing filmthickness (22, 28, and 56 mm) on tin-plated steel, and decayed at a muchslower rate for the specimen with the thickest film (Clausen, 1993).

Nominal initial wet-film thicknesses for paints applied to gypsum boardusing a roller have been reported to be approximately 100 µm (Chang et al.,1997; Roache et al. 1996, Sparks et al., 1999b). Clausen (1993) and Clausenet al. (1991) reported final dry film thicknesses of between 22 and 63 µmusing a Twintector Elcometer device for latex paint applied to tin-platedsteel.

In the future, it might be possible to gain additional insights into theemissions process by varying the thickness of paint applied to gypsum boardand measuring temporal differences in both emissions and mass recovery ofindividual components. For example, one would not expect significant dif-ferences in emissions from the wet film if the film composition is relatively

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

1068 R. L. Corsi and C.-C. Lin

homogeneous unless a greater fraction of film mass is adsorbed by thesubstrate for thinner films. Similarly, a lower long-term recovery from appli-cations involving thick films might suggest the importance of the dry-filmas a barrier for diffusion of TMPD-MIB and other components out of thesubstrate and film itself.

Effects of Paint Composition on TMPD-MIB Emissions

The author was unable to identify any published research that explicitly ad-dresses the effects of paint composition on TMPD-MIB emissions. However,it is reasonable to speculate on several compositional variables that mayaffect TMPD-MIB emissions. For example, for an otherwise equivalent paintcomposition, an increase in TMPD-MIB weight fraction should lead to anincrease in TMPD-MIB emissions, presumably during both paint drying andthereafter.

The effects of variations in other paint components, particularly thesolvents ethylene glycol and propylene glycol, may also affect TMPD-MIBemissions. As described elsewhere in this paper, increases in EG and PGweight fractions may increase the solubility of TMPD-MIB in the liquid phase,thereby increasing the TMPD-MIB mass available for volatilization duringpaint drying. Increasing values of EG and PG content may also influencethe nature and extent of TMPD-MIB adsorption within a substrate throughmulti-component competition for sorption sites. However, the importanceof these effects cannot be gleaned from previous publications related toemissions from paint products, and is an area in need of additional research.

After paint has dried, long-term emissions of TMPD-MIB are presumablydependent, at least in part, on the nature of TMPD-MIB molecular diffusionthrough and/or out of the dried paint film. The nature of a dried paint filmon the effective diffusion coefficient of TMPD-MIB may vary depending onthe solids composition of a paint mixture. For example, flat latex paintstend to have greater amounts of “filler” material relative to semi-gloss latexpaints that often contain a much larger fraction of acrylic resins. Thus, theeffective diffusion coefficient through the dry film of flat latex paint may beconsiderably different than that through semi-gloss latex paint. Additionalresearch is needed to assess the significance of differences in dry paint filmproperties on TMPD-MIB emissions.

Effects of Environmental Conditions on TMPD-MIB Emissions

Several researchers have studied the effects of different environmental con-ditions on (S)VOC emissions from latex paint. The primary variables thathave been studied are air temperature, relative humidity (RH), and air speedabove the painted surface.

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

Emissions of TMPD-MIB from Latex Paint 1069

Wolkoff (1998) applied a latex paint to gypsum board specimens, al-lowed the films to set for 24 hours, and then used a Field and LaboratoryEmission Cell (FLEC)©R to measure the concentration-time profiles of 1,2-propanediol and TMPD-MIB over 250 days. The effects of temperatures of23, 35, and 60◦C on (S)VOC emissions were assessed. For TMPD-MIB, in-creased temperatures led to corresponding elevations of the concentration-time profile for the first week, with no difference after the first week. For1,2-propanediol at 60◦C, the concentration-time profile dropped below thatobserved at lower temperatures after just two days, and approached 0 µg/m3

after only one week.Fang et al. (1999) applied a latex paint to nine gypsum board spec-

imens with subsequent conditioning in CLIMPAQ chambers. Area-specificemission rates were measured after three weeks. Three temperatures (18,23, and 28◦C) were used for each of three values of relative humidity (30,50, and 70%). There was no clear trend in emission rate for TMPD-MIB withincreasing temperature at 30% and 50% RH. However, a significant increasein emissions of TMPD-MIB was observed with increasing temperature at theelevated relative humidity (70%). A four-fold increase in emission rate (∼51to 216 µg/m2-hr) was observed with a 10◦C increase in temperature (18 to28◦C) at 70% RH.

Van der wal et al. (1997) applied a latex paint to tin-plated steel andmeasured concentrations of TMPD-MIB over a two-week period in two smalllaboratory chambers. One chamber was operated at 23◦C and the other at30◦C. The relative humidity was maintained at 45% in each chamber. Thepeak concentration of TMPD-MIB was approximately twice as great at theelevated temperature, followed by a more rapid decay rate.

Published results are mixed with respect to the effects of relative hu-midity on (S)VOC emissions from latex paint. While an increase from 0% to50% RH had a significant impact on emissions of 1,2-propanediol (higheremissions at higher RH), there was no observable effect on emissions ofTMPD-MIB over 25 days. This result is consistent with findings by Hodg-son (1999), who observed little effects of changes in RH on emissions ofTMPD-MIB from a latex primer sealer and a flat latex paint. Roache et al.(1996) also observed little effect of variations in RH between 24% and 79%on emissions of either TMPD-MIB or BEE, but did observe significant re-ductions in EG emissions with increasing RH. In contrast, Fang et al. (1999)reported an increase up to a factor of four in emissions of TMPD-MIB asrelative humidity was varied from 30% to 70% at temperatures above 23◦C.A less significant increase was observed at 18◦C. Finally, TMPD-MIB concen-trations were positively correlated with RH in school classrooms in Sweden(Kim et al., 2007).

Variations in air speed above a painted surface do not appear to havea significant influence on TMPD-MIB from latex paint. Wolkoff (1998) var-ied air speed between 1–9 cm/s above latex-painted gypsum board, values

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

1070 R. L. Corsi and C.-C. Lin

consistent with those measured by Sparks et al. (1999b) near the walls of aUSEPA test house. He observed little differences in emissions of TMPD-MIBover this range of air speeds. This result is consistent with Hodgson andShimer (1999), who observed little effect of increasing air exchange rate ormixing intensity of air (additional fans) on the concentration of TMPD-MIBduring large chamber experiments.

Effects of Sorptive Sinks on Fate and Persistence of TMPD-MIB

Sorptive interactions with chamber or indoor materials after being emittedfrom paint may affect determination of emission rates of TMPD-MIB, as wellas its persistence within building environments. For example, Clausen et al.(1991) noted potential problems with TMPD-MIB adsorption and prolongeddesorption from the walls of small chambers.

Chang et al. (1998) exposed 17-year-old painted gypsum board removedfrom a USEPA test house to relatively high gaseous concentrations of TMPD-MIB, BEE, EG, and PG, followed by a 300-hour purge/desorption stage. Theydetermined a mass recovery of only 18% for TMPD-MIB. Recoveries for BEE,EG, and PG were somewhat lower, a trend consistent with other studiesinvolving latex paint application to gypsum board (e.g., Chang et al., 1997).

Sparks et al. (1999a) completed small chamber experiments to deter-mine adsorption and desorption rate constants for gypsum board and carpetexposed to TMPD-MIB. The rate constants were determined based on theclassic linear surface sorption model:

Rsin k = kaC − kdMsin k (1)

where, Rsink = the rate of removal of a chemical to a material (mg/m2-hr),ka = an adsorption rate constant (m/hr), kd = a desorption rate constant(1/hr), C = the concentration of the chemical of interest in the gas-phase(chamber or room concentration) (mg/m3), and Msink = the concentration ofthe chemical of interest adsorbed to the sink material (mg/m2). The equilib-rium partition coefficient (Keq) for each material was determined as the ratioof the adsorption and desorption rate constants. Relevant sorption parame-ters for TMPD-MIB are summarized in Table 5. Note that for gypsum board,the adsorption and desorption rate constants and equilibrium partition coef-ficient for TMPD-MIB were reasonably similar to those reported for EG, PG,and BEE. However, ka was lower, and kd was greater, for TMPD-MIB relativeto EG, PG, and BEE for carpet. The resulting Keq is an approximate orderof magnitude lower for TMPD-MIB than for the other components of latexpaint. This suggests that the strong polar-polar interactions between gyp-sum board and each polar compound effectively dampens any differencesin sorption parameters between the compounds, while the same is not truefor the less polar surfaces of carpet components. Interestingly, the results

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

Emissions of TMPD-MIB from Latex Paint 1071

TABLE 5. Parameters for sorptive interactions between TMPD-MIB and gypsum board/carpet.

Parameter in Eq. (1) Gypsum board Carpet

ka(m/hr) 1.76 ± 0.17 0.84 ± 0.083kd (1/hr) 0.0048 ± 0.00075 0.016 ± 0.0023Keq (m) 369 52

From Sparks et al. (1999a).Values of ka and kd for use in Eq. (1). Keq based on ka/kd (mean values of each). ±based on onestandard deviation.

of Won et al. (2000) suggest that because TMPD-MIB has a lower vaporpressure than the other constituents of latex paint, its equilibrium partitioncoefficient should be greater than those of the other compounds for carpet.The opposite was noted by Sparks et al. (1999a).

Effects of Chemical Reactions on TMPD-MIB Emissions

The authors could find no published evidence of reduction/oxidation chem-istry affecting the emissions of TMPD-MIB from paint or its fate in indoorenvironments. However, there is some evidence that at least one of the iso-mers of TMPD-MIB may undergo hydrolysis reactions over time scales rele-vant to emission time scales in indoor environments. For example, Shieldsand Weschler (1992) measured the concentrations of each TMPD-MIB isomerover the course of three months after latex paint applications in a telephoneswitching station. After each of two painting events, they observed a greaterrelative reduction in the x = 1 isomer than the x = 3 isomer in room air.They speculated that the greater reduction in concentration of the x = 1 iso-mer was due to hydrolysis, as the ester of a primary alcohol (x = 1 isomer)tends to hydrolyze more rapidly than the ester of a secondary alcohol (x =3 isomer). If so, these results are consistent with unpublished findings thatindicate measurable first-order hydrolysis rates for TMPD-MIB in alkaline so-lutions (Roser, 1992). However, additional research is needed, particularlyover the time scales of months to years, to identify whether chemical reac-tions play a role in reduction of TMPD-MIB source strengths, as well as theultimate fate of TMPD-MIB following the application of latex paints.

Models for Estimation of TMPD-MIB Emissions

The processes that affect TMPD-MIB emissions are clearly dynamic and com-plex, and are not fully understood. As such, purely mechanistic models forthe emissions process have yet to be developed, and data to support ahighly mechanistic model of the emissions process do not exist in the pub-lished literature. Nevertheless, several researchers have developed empiricaland semi-empirical models to estimate species emissions following paint

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

1072 R. L. Corsi and C.-C. Lin

applications, and have reported model parameters for TMPD-MIB. Thosemodels are described in this section, along with several salient features asso-ciated with their development and parameters relevant to TMPD-MIB, whereappropriate. The reader is referred to the original publications for detailsrelated to model derivations and experimental methodologies used to deter-mine model parameters.

A simple first-order decay equation has been used to estimate VOCemissions from architectural coatings (e.g., Van der wal et al., 1997):

Ea(t) = M(0)ke−kt (2)

where M(0) = the initial (t = 0) amount of chemical applied per unit area(mg/m2), and k = a first-order decay constant (1/hr). (All other variablesare as defined above.) The empirical nature of this model precludes its ef-fective application outside of the conditions for which the decay constantis determined experimentally. A significant problem with Eq. (2) is that ifthe decay constant is based on a limited amount of short-term data, themodel trends toward underestimation of long-term emissions. Eq. (2) is inca-pable of effectively predicting emissions during both the paint drying phase(evaporation-controlled phase) and dried paint phase (diffusion-controlledphase).

To overcome problems associated with the prediction of long-term emis-sions, some researchers have employed a double-exponential decay modelto estimate VOC emissions from architectural coatings (e.g., Chang et al.,1997; Wilkes et al., 1996):

Ea(t) = E1a(0)e−k1t + E2a(0)e−k2t (3)

where E1a(0) and E2a(0) = the initial (t = 0) area-specific emission factorsfor phases 1 and 2, respectively (mg/m2-hr), k1 and k2 = the emission decayrate constants for phases 1 and 2, respectively (1/hr), and all other variablesare as defined above. In this model, short-term emissions are meant to beaccounted for by the first term on the right-hand-side of Eq. (3). Longer-term emissions are meant to be accounted for by the second term on theright-hand-side of the equation. As such, k1 � k2.

While Eq. (3) can provide better estimates than Eq. (2) for long-termemissions, it still suffers from the limitations described above in terms ofits empirical nature (i.e., it cannot effectively be used to estimate emissionsfor conditions other than those used to experimentally determine k1 andk2). Wilkes et al. (1996) suggested the possibility of overcoming part of thislimitation by correlating k1 to chemical vapor pressure and k2 to molecu-lar weight of the emitted component. However, this approach still fails toaccount for the changing composition of a wet paint film (changing molefractions of emitting constituents), the nature of the dried paint film (whichvaries in time and between paints), and environmental conditions.

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

Emissions of TMPD-MIB from Latex Paint 1073

Clausen et al. (1991) and Clausen (1993) presented Eqs. (4) and (5) topredict emissions following latex paint applications:

Ea(t) = M(0)

(kε1

L

)e

−(

kε1tL

)(4)

Ea(t) = M(0)

(kD1

L2

)e

−(

kD1t

L2

)(5)

where kε1 = the rate constant for evaporation controlled emissions forL = 1 µm (µm/hr), kD1 = the rate constant for diffusion-controlled emis-sions for L = 1 µm (µm2/hr), and L = effective thickness of source (µm). (Allother variables are as defined previously.) Eq. (4) is intended for evaporation-controlled emissions (i.e., it is applicable when diffusion through the bound-ary layer above the paint film is much slower than through the dry paint filmitself). It is based on application of Fick’s first law with an assumed linearconcentration gradient through a laminar boundary layer above the paint.The rate constant for evaporation-controlled emissions includes the diffusioncoefficient in air for the chemical of interest as well as the thickness of thelaminar boundary layer. Eq. (5) is intended for diffusion-controlled emissions(i.e., when diffusion within the dry paint film is slower than through the lam-inar boundary layer above the paint). It is based on an application of Fick’ssecond law, with an assumed zero concentration for the chemical of interestat the film surface. It is further assumed that there is a zero flux (no fluxcondition) at the interface of the paint film and underlying substrate. Whilethis assumption might be valid for impermeable surfaces such as aluminumand glass, it is likely not valid for porous materials such as gypsum board orconcrete.

While Eqs. (4) and (5) are more mechanistic in derivation than Eqs. (2)and (3), they reduce to dual application of Eq. (3), for which kε1/L = k1 andkD1/L2 = k2. Furthermore, as with previous models, application of Eqs. (4)and (5) are relevant only to those conditions for which model parameters,particularly kε1 and kD1, are determined.

Guo et al. (1996) proposed the following model to account for bothshort-term and long-term chemical emissions:

Ea(t) = MV(0)ke−kt + αfDMD(0)e−2fDt1/2

t1/2(6)

where MV (0) = initial (t = 0) mass of chemical available for emissions byevaporation (mg/m2), Md(0) = initial mass of chemical available for emis-sions by molecular diffusion (mg/m2), and Ea(t), k, and t are as definedpreviously. The term α is an adjusting factor that is needed to account forthe fact that diffusion-controlled emissions cannot become important untilthe paint film is dried. Guo et al. (1996) selected the following equation

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

1074 R. L. Corsi and C.-C. Lin

for α:

α = (1 − e−kt)2 (7)

The term fd is referred to as a diffusion constant, which is actually a pa-rameter defined by Eq. (8) and stems from the solution of a partial differentialequation to represent the time and one-dimensional space-distribution of achemical diffusing through a layer of thickness λ, as described by Treybal(1968):

fD = 0.632D1/2

λ(8)

where D = the effective diffusion coefficient through some medium (m2/hr)and λ = the thickness of the diffusion later (m).

Sparks et al. (1999a, 1999b) developed and applied what is arguablythe most mechanistic of models published for estimation of chemical emis-sions following paint applications. The model is effectively a coupling of thefollowing three equations:

Ea(t) = km

(CVp

MV(t)

MV(0)− C(t)

)+

(1 − MV(t)

MV(0)

)2 fDMD(t)

t1/2(9)

dMV(t)

dt= −km

(CVp

MV(t)

MV(0)− C(t)

)(10)

dMD(t)

dt= −

(1 − MV(t)

MV(0)

)2 fDMD(t)

t1/2(11)

where km = gas-phase mass transfer coefficient (m/hr), CVp = vapor pressurefor chemical of interest in concentration units, i.e., saturation concentration(mg/m3), C(t) = concentration in room air (mg/m3), MV(t) = mass of chemi-cal available for emissions by evaporation at time t (mg/m2), MV(0) = initial(t = 0) mass of chemical available for emissions by evaporation (mg/m2)as described above for Eq.(6), and Md(t) = mass of chemical available foremissions by molecular diffusion (mg/m2). (All other variables are as definedabove.) The gas-phase mass transfer coefficient (km) can be estimated basedon Nusselt and Reynolds numbers as described by Sparks et al. (1999b).

Model parameters relevant to TMPD-MIB emissions are listed in Table 6.To the extent possible, paint and substrate properties, as well as experimen-tal conditions for which model parameters were derived, are also listed inTable 6. The small number of studies for which TMPD-MIB parameters areavailable for a common substrate, and the different conditions for whichmodel parameters were derived, precludes a detailed comparison of modelpredictions.

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

Emissions of TMPD-MIB from Latex Paint 1075

TABLE 6. Reported model parameters for TMPD-MIB emissions.

EquationsPaint/substrate

propertiesExperimentalconditions Model parameters Reference

2 Al plates. Acrylic paint:composition notprovided.

6.3 L chamber(0.95 ACH)24-hourdryingperiod priorto measure-ments.T = 23◦C &30◦C; 45%RH

M(0) (mg/m2)23◦C/30◦C

·Isomer 1 = 64/41·Isomer 2 = 101/66

k (hr−1× 10−3)·Isomer 1 = 59/230·Isomer 2 = 41/260

Van der walet al.(1997)

3 Stainless steel andgypsum board. Paint:Nominal wet film =100 µm.TVOC = 45.4 mg/g;PG = 2.32; EG = 24.0;BEE = 4.98;TMPD-MIB = 13.5;diethylene glycol =0.59

53 L chamber(0.5 ACH)

23◦C; 50% RH

Stainless steel(short-term)

·E1a(0) = 30 mg/m2-hr·E2a(0) = 0 mg/m2-hr·k1 = 0.0169/hr

Gypsum board·E1a(0) = 29.7mg/m2-hr

·E2a(0) = 15.9mg/m2-hr

·k1 = 0.795/hr·k2 = 0.0317/hr

Chang et al.(1997)

3 Pre-painted gypsumboard.

Latex paint: EG (54%),PG, BEE, TMPD-MIB(30%)

Smallchambers

k1 = 0.944/hrk2 = 0.013/hrE1a(0) = 1.46 mg/hrE2a(0) = 0.22 mg/hr

Wilkes et al.(1996)

2, 4, 5 Tin-plated stainlesssteel.

234 L chamber(0.25 ACH)

Values for L = 22/28/56µm

Clausen(1993)

Waterborne paint =white spirit;1,2-propanediol, BEE;TMPD-MIB (fractionsnot provided). Dryfilm thickness = 22,28, 56 µm.

23 ± 0.6◦C; 45± 3% RH

·M(0) (mg/m2): 35/ 46/410

·k (hr−1): 0.015/ 0.016/0.0048

·kε1 (µm/hr): 0.33/ 0.46/0.27

·kD1 (µm2/hr): 7/ 13 /156, 7 Gypsum board. Paint:

4.5% TVOC: EG =53%; TMPD-MIB =30%; BEE = 11%;PG = 5%; diethyleneglycol = 1%

53 L chamber(0.5 ACH)

23◦C; 50% RH

k = 0.0635 /hrMV(0) = 404 mg/m2

Md(0) = 1,465 mg/m2

fd = 0.00173 /hr1/2

Guo et al.(1996)

9–11 Gypsum board. Paint:TVOC = 45 mg/g;EG = 24; TMPD-MIB = 13; BEE = 5;PG = 2; diethyleneglycol = 2

53-L chamber(0.5 ACH)

23◦C; 50% RH

CVO = 20 g/m3

MV(0) = 496 mg/m2

fd = 0.0012/hr1/2

Sparks et al.(1999a)

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

1076 R. L. Corsi and C.-C. Lin

SUMMARY AND RESEARCH NEEDS

The existing literature related to TMPD-MIB emissions from latex paint issparse and generally insufficient to make highly certain conclusions regard-ing the nature of the emissions process or the ultimate fate of TMPD-MIB. Thesame is true for other volatile components of latex paint, and the small baseof published data related to TMPD-MIB precludes any attempts to extrapolateinformation to the other compounds. Nevertheless, the following observa-tions, conclusions, and recommendations for future research are made basedon this review:

� Published measurements of TMPD-MIB emissions following latex paintapplications are generally limited by a lack of analysis of initial paint com-position, as well as numerous reports of paint applications to impermeablesubstrates such as stainless steel and glass.

� Long-term emissions data for TMPD-MIB are sparse, but existing data dosuggest continued emissions for up to at least a year after paint application.

� The range of TMPD-MIB recoveries in air following paint applicationsto gypsum board is relatively large, with differences being difficult toascertain based on available data.

� The substrate to which paint is applied can have a significant influenceon (S)VOC emissions, including TMPD-MIB. This is particularly true whencomparing emissions from impermeable materials (e.g., stainless steel)with emissions from porous materials such as gypsum board or concrete.

� It is clear that either liquid-solid and/or gas-solid sorptive interactionsbetween TMPD-MIB and gypsum board retard short-term emissions andprolong the emissions process relative to emissions from impermeablesubstrates.

� The most important factor associated with paint application appears to bethe thickness of the paint film. However, additional work is needed toassess the effects of paint composition on TMPD-MIB emissions.

� The effects of temperature and relative humidity on TMPD-MIB emissionsfrom latex paint are not well understood and may be coupled. It does notappear that air speed above painted surfaces has a significant influenceon TMPD-MIB emissions.

� The migration and storage of TMPD-MIB into gypsum board is observedto be significant; absolute mass storage in the substrate is typically greaterthan that in the dry paint film after six months or more.

There is a clear need for long-term mass recovery data associated withTMPD-MIB following applications to wood products, concrete, and othercementatious materials.

Significantly more research, as well as new research approaches, isneeded to improve the existing knowledge base associated with TMPD-MIB

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

Emissions of TMPD-MIB from Latex Paint 1077

emissions from latex paint and to ultimately allow for the development ofimproved mechanistic models of both short-term and long-term emissions.

ACKNOWLEDGMENTS

The authors wish to thank the Eastman Chemical Company for an unre-stricted gift to The University of Texas that made this work possible. Specialthanks are given to Dr. Bruce Gustafson of Eastman Chemical and Dr. DavidMorgott, as well as to Mr. Robert Avery for their expeditious provision of in-formation when requested. During this work, Chi Chi Lin was supported byGrant No. T42CCT610417 from the Southwest Center for Occupational andEnvironment Health, as supported by the National Institute for Occupationaland Environmental Health (NIOSH)/Centers for Disease Control and Preven-tion (CDC). The corresponding author wishes to thank the late Roxanne A.Shepherd for her dedicated support, love, and inspiration.

REFERENCES

Agency for Toxic Substances and Disease Registry (ATSDR). (1997). Toxi-cological profile for ethylene glycol and propylene glycol. Available at:http://www.atsdr.cdc.gov/toxprofiles/tp96.html.

Avery, R. (2004). Personal communication, Eastman Chemical Company.Carter, W.P.L., and Malkina, I.L. (2005). Evaluation of atmospheric impacts of selected

coatings VOC emissions. Report to the California Air Resources Board, contractno. 00-333.

Censullo, A.C., Jones, D.R., and Wills, M.T. (1996). Improvement of speciation pro-files for architectural and industrial maintenance coating operations. Report tothe California Air Resources Board, contract no. 93–319.

Chang, J.C.S., Sparks, L.E., Guo, Z., and Fortmann, R. (1998). Evaluation of sinkeffects on VOCs from latex paint. Journal of the Air & Waste ManagementAssociation, 48, 953–958.

Chang, J.C.S., Tichenor, B.A., Guo, Z., and Krebs, K.A. (1997). Substrate effects onVOC emissions from a latex paint. Indoor Air, 7, 241–247.

Clausen, P. (1993). Emission of volatile and semi-volatile organic compoundsfrom waterborne paints—the effect of the film thickness. Indoor Air, 3, 269–275.

Clausen, P.A., Wolkoff, P., Holst, E., and Nielsen, P.A. (1991). Long-term emission ofvolatile organic compounds from waterborne paints—methods of comparison.Indoor Air, 4, 562–576.

De Bortoli, M., Kefalopoulos, S., Kirchner, S., Schauenburg, H., and Vissers, H.(1999). State-of-the-art in the measurement of volatile organic compounds emit-ted from building products: Results of European interlaboratory comparison.Indoor Air, 9, 103–116.

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

1078 R. L. Corsi and C.-C. Lin

Eastman Chemical Company. (2003). Material safety data sheet: Texanol©Rester

alcohol. MSDSUSA/ANSI/EN/150000000148/version 5.0. Available at:http://www.eastman.com (revised 1 August 2003).

Fang, L., Clausen, G., and Fanger, P.O. (1999). Impact of temperature and humidityon chemical sensory emissions from building materials. Indoor Air, 9, 193–201.

Fortmann, R.C., Sheldon, L.S., Keever, J.T., Whitakerand, D.A. (1993). Comparisonof methods for analyzing emissions from architectural coatings used indoors.Proceedings of Indoor Air ‘93, 2, 263–268.

Gehrig, R., Hill, M., Zellweger, C., and Hofer, P. (1993). VOC emissions from wallpaints—a test chamber study. Proceedings of Indoor Air ‘93, 2, 431–436.

Girman, J.R., Hadwen, G.E., Burton, L.E., Womble, S.E., and McCarthy, J.F. (1999). In-dividual volatile organic compound prevalence and concentrations in 56 build-ings of the building assessment survey and evaluation (BASE) study. Proceedingsof Indoor Air ‘99, 2, 460–465.

Guo, Z., Fortmann, R.C., Marfiak, S., Tichenor, B., Sparks, L., Chang, J., and Mason,M. (1996). Modeling the VOC emissions from interior latex paint applied togypsum board. Proceedings of Indoor Air ‘96, 1, pp. 987–992.

Hodgson, A.T.(1999). Common indoor sources of volatile organic compounds: Emis-sion rates and techniques for reducing consumer exposures. California Air Re-sources Board Report, Contract No. 95-302. E.O. Lawrence Berkeley NationalLaboratory, Sacramento, Calif.

Hodgson, A.T., Rudd, A.F., Beal, D., and Chandra, S. (2000). Volatile organic com-pound concentrations and emission rates in new manufactured and site-builthouses. Indoor Air, 10, 178–192.

Hodgson, A.T., and Shimer, D.A. (1999). Techniques for reducing exposures tovolatile organic compounds associated with new construction and renovation.Proceedings of Indoor Air ‘99, 4, 622–627.

Kim, J.L., Elfman, L., Mi, Y., Wieslander, G., Smedje, G., and Norback, D. (2007).Indoor molds, bacteria, microbial volatile organic compounds and plasticizers inschools—associations with asthma and respiratory symptoms in pupils. IndoorAir, 17, 153–163.

Klimont, Z., Streets, D.G., Gupta, S., Cofala, J., Fu, L.X., and Ichikawa, Y. (2002).Anthropogenic emissions of non-methane volatile organic compounds in China.Atmospheric Environment, 36(8), 1309–1322.

Knudsen, H.N., Kjaer, U.D., Nielson, P.A., and Wolkoff, P. (1999). Sensory andchemical characterization of VOC emissions from building products: Im-pact of concentration and air velocity. Atmospheric Environment, 33, 1217–1230.

Krebs, K., Lao, H.C., Fortmann, R., and Tichenor, B. (1995). Test methods for de-termining short and long term VOC emissions from latex paint. EngineeringSolutions to Indoor Air Quality Problems, a specialty conference sponsored bythe Air & Waste Management Association and the United States EnvironmentalProtection Agency, pp. 71–75.

Lawrimore, J.H., and Aneja, V.P. (1997). A chemical mass balance analysis of non-methane hydrocarbon emissions in North Carolina. Chemosphere, 35(11), 2751–2765.

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

Emissions of TMPD-MIB from Latex Paint 1079

Lin, C.C., and Corsi, R.L. (2007). Texanol©R

ester alcohol emissions from latex paints:Temporal variations and multi-component recoveries. Atmospheric Environ-ment, 41(15), 3225–3234.

Nielsen, G.D., Hansen, L.F., and Wolkoff, P. (1997). Chemical and biological evalua-tion of building material emissions, II. Approaches for setting indoor standardsor guidelines for chemicals. Indoor Air, 7, 17–32.

Norback, D., Wieslander, G., and Edling, C. (1995). Occupational exposure to volatileorganic compounds (VOCs), and other air pollutants from the indoor applicationof water-based paints. Annals of Occupational Hygiene, 39(6), 783–794.

Rector, D. (1995). Personal communication, Eastman Chemical.Roache, N., Howard, E., Guo, Z., and Fortmann, R. (1996). Observations on applica-

tion of the field and laboratory emission cell (FLEC) for latex paint emissions—effect of relative humidity. Proceedings of Indoor Air ‘96, 2, pp. 657–662.

Roser, K.S. (1992). Abiotic degradation: Hydrolysis as a function of pH—Texanolester alcohol. Internal report, Eastman Kodak Company.

Sheldon, L.S., and Naugle, D.F. (1994). Determination of test methods for inte-rior architectural coatings. Revised final report to ICF Inc., USEPA Contact No.68-D2–0131. Report RTI/5522/042-02 FR. Research Triangle Institute, ResearchTriangle Park, North Carolina.

Shields, H.C., and Weschler, C.J. (1992). Volatile organic compounds measured ata telephone switching center from 5/30/85-12/6/88: A detailed cased study.Journal of the Air & Waste Management Association, 42, 792–804.

Silva, G.V., Vasconcelos, M.T.S.D., Santos, A.M., and Fernandes, E.O. (2003). Com-parison of the substrate effects on VOC emissions from water based varnish andlatex paint. Environmental Science and Pollution Research, 10(4), 209–216.

Sparks, L.E., Guo, G., Chang, J.C., and Tichenor, B.A. (1999a). Volatile organiccompound emissions from latex paint, part 1: Chamber experiments and sourcemodels. Indoor Air, 9, 10–17.

Sparks, L.E., Guo, Z., Chang, J.C.S., and Tichenor, B.A. (1999b). Volatile organiccompound emissions from latex paint, part 2: Test house studies and indoor air(IAQ) modeling. Indoor Air, 9, 18–25.

Tichenor, B.A. (1995). Evaluation of emissions from latex paint. Paper presented atthe Low- and No-VOC Coating Technologies 2nd Biennial International Confer-ence, Durham, North Carolina.

Tichenor, B.A., and Sparks, L.E. (1996). Managing exposure to indoor air pollutantsin residential and office environments. Indoor Air, 6, 259–270.

Treybal, R.E. (1968). Mass transfer operations. New York, McGraw-Hill, Inc.United States Department of Commerce. (2003). Paint and allied products: 2002.

Economics and Statistics Administration, U.S. Census Bureau, MA325F(02)-1.United States Environmental Protection Agency. (2002). A small chamber test method

for the measurement of volatile organic compounds and hazardous air pollutantsfrom alkyd and latex paints. Office of Research and Development, ResearchTriangle Park, North Carolina. Inside IAQ, Fall/Winter, pp. 1, 3–7.

Van der wal, J.F., Hoogeveen, A.W., and Wouda, P. (1997). The influence of temper-ature on the emission of volatile organic compounds from PVC flooring, carpet,and paint. Indoor Air, 7, 215–221.

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14

1080 R. L. Corsi and C.-C. Lin

Verschueren, K. (1996). Handbook of environmental data on organic chemicals. 3rded. Van Nostrand Reinhold, New York.

Wadden, R.A., Scheff, P.A., and Uno, I. (1994). Receptor modeling of VOCs-2. devel-opment of VOC control functions for ambient ozone. Atmospheric Environment,28(15), 2507–2521.

Wilkes, C., Koontz, M., Ryan, M., and Cinalli, C. (1996). Estimation of emissionprofiles for interior latex paints. Proceedings of Indoor Air ‘96, 2, 55–60.

Wolkoff, P. (1998). Impact of air velocity, temperature, humidity, and air onlong-term VOC emissions from building products. Atmospheric Environment,32(14/15), 2659–2668.

Won, D., Corsi, R.L., and Rynes, M. (2000). New indoor carpet as an adsorptivereservoir for volatile organic compounds. Environmental Science & Technology,34(19), 4193–4198.

Yu, C.W.F., and Crump, D.R. (1999). Factors influencing the measurement of VOCemission from paints for interior application using test chambers. Proceedingsof Indoor Air ‘99, 5, 155–160.

Ziemer, P.D., Woo, J., and Anagnostou, T. (2000). Study of odor qualification ofsolvents used in coating compositions. Journal of Coatings Technology, 72, 97–102.

Dow

nloa

ded

by [

Nor

th D

akot

a St

ate

Uni

vers

ity]

at 1

4:24

07

Nov

embe

r 20

14