CREST 2009 Texanol Review

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 30

This article was downloaded by: [Corsi, Richard L.

]
On: 4 December 2009
Access details: Access Details: [subscription number 917383724]
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.informaworld.com/smpp/title~content=t713606375

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

Online publication date: 02 December 2009

To cite this Article CORSI, RICHARD L. and LIN, CHI-CHI(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
To link to this Article: DOI: 10.1080/10643380801977925
URL: http://dx.doi.org/10.1080/10643380801977925

PLEASE SCROLL DOWN FOR ARTICLE

Full terms and conditions of use: http://www.informaworld.com/terms-and-conditions-of-access.pdf

This article may be used for research, teaching and private study purposes. Any substantial or
systematic reproduction, re-distribution, re-selling, loan or sub-licensing, systematic supply or
distribution in any form to anyone is expressly forbidden.

The publisher does not give any warranty express or implied or make any representation that the contents
will be complete or accurate or up to date. The accuracy of any instructions, formulae and drug doses
should be independently verified with primary sources. The publisher shall not be liable for any loss,
actions, claims, proceedings, demand or costs or damages whatsoever or howsoever caused arising directly
or indirectly in connection with or arising out of the use of this material.
Critical Reviews in Environmental Science and Technology, 39:1052–1080, 2009
Copyright © Taylor & Francis Group, LLC
ISSN: 1064-3389 print / 1547-6537 online
DOI: 10.1080/10643380801977925

Emissions of 2,2,4-Trimethyl-1,3-Pentanediol
Monoisobutyrate (TMPD-MIB) from Latex Paint:
A Critical Review

RICHARD L. CORSI1 and CHI-CHI LIN2


1
Department of Civil, Architectural and Environmental Engineering, The University of Texas,
Austin, Texas, USA
2
Department of Civil and Environmental Engineering, National University of Kaohsiung,
Kaohsiung City, Taiwan
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

The significance of latex paint as a source of indoor volatile or-


ganic compounds is underscored by the large volume produced
for 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 on
measurements of TMPD-MIB emissions, experimental recoveries of
TMPD-MIB, and factors that affect the emissions process. Published
models that attempt to describe TMPD-MIB emissions following latex
paint applications are summarized. Finally, a critical assessment
of the state of knowledge related to TMPD-MIB emissions from latex
paints is presented, along with a summary of continuing research
needs.

R
KEY WORDS: Texanol , fate, architectural coatings, gypsum
board, models

INTRODUCTION

Architectural coatings applied to building materials (substrates) are a source


of emissions of volatile and semi-volatile organic compounds ((S)VOCs) to
both indoor and outdoor atmospheres. Concerns related to such emissions
are 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, TX
78712, USA; Tel.: (512) 475-8617; Fax: (512) 471-1720; E-mail: corsi@mail.utexas.edu

1052
Emissions of TMPD-MIB from Latex Paint 1053

reactive and can contribute to the production of photochemical smog in


urban air sheds (Klimont et al., 2002; Lawrimore & Aneja, 1997; Wadden
et al., 1994). As such, the inclusion of speciated (S)VOC emissions from
architectural 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 periods
of 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 covered
by indoor paints, primarily water-based paints, and the potential impacts on
indoor 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 795
million gallons of architectural coatings were produced in the United States
in 2005 (United States Department of Commerce, 2006). Of this, 65% (519
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

million gallons) was classified as being for interior use. Of the interior-type
coatings, 89% (460 million gallons) was classified as interior water-type, as
opposed 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 flat
water-thinned paints/tinting bases and semi-gloss, eggshell, satin, and other
water-thinned paints/tinting bases. However, despite the large amounts of
architectural coatings now being produced and used around the world, as
well as the potential impacts of emissions from architectural coatings on out-
door and indoor air quality, there is still a great deal of uncertainty regarding
both the impacts of such emissions and the nature of emissions processes.
This review focuses on one important component of latex paints, 2,2,4-

R
trimethyl-1,3-pentanediol monoisobutyrate, often referred to as Texanol es-
ter alcohol and hereafter denoted TMPD-MIB. There are several compelling
reasons to review the existing knowledge base related to TMPD-MIB, the
first 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. The
global production rate of TMPD-MIB is estimated at 98,000 metric tons/year
(Rector, 2005). Further, TMPD-MIB has a low-odor threshold concentration
and is thus often detected for hours to days after a painting event has oc-
curred. It has an airway irritation threshold concentration that is reasonably
achieved 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 recent
study indicated a positive association between TMPD-MIB concentrations
in school classrooms and nocturnal breathlessness in the home (Kim et al.,
2007). However, the home environment was not characterized in the study.
1054 R. L. Corsi and C.-C. Lin

TMPD-MIB has come under some scrutiny as a precursor to urban ozone


formation. 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 the
fate 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 of
other oxygenated components used in architectural coatings and cleaners
and 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 paints
and potential improvements in the existing knowledge-base related to such
emissions. Past experimental research is described, with an emphasis on
measurements of TMPD-MIB emissions following latex paint applications,
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

experimental recoveries of TMPD-MIB, and factors that affect the emissions


process. Published models that attempt to describe TMPD-MIB emissions fol-
lowing latex paint applications are summarized. Finally, a critical assessment
of the state of knowledge related to TMPD-MIB emissions from latex paints
is presented, along with a summary of continuing research needs.

BACKGROUND
Latex Paints
Latex paints are water-thinned paints that contain significantly lower amounts
of VOCs than oil-based paints, and tend to dry much more rapidly. Inter-
estingly, water-based paints have been used since pre-industrial times. For
example, 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 fine
particles that are used to provide color and opacity or covering power. The
particles are generally inorganic in nature (e.g., titanium dioxide). Binders
assist with binding of pigments and additives together and provide adhesion
to the substrate being coated. Higher performance latex paints often include
acrylic resins resulting from the polymerization of derivatives of acrylic acids
as a binder. However, other binders such as polyvinyl acetate are also em-
ployed in latex paints. Solvents aid in maintaining unused paint in a liquid
form prior to application and allow pigment and binder solids to behave
as a fluid during application. Solvents are intended to evaporate completely
following application, with evaporation rates largely dependent on the na-
ture of the solvent mixture and indoor environmental conditions (e.g., air
speed, humidity, temperature). Water and glycol solvents, principally ethy-
lene glycol and propylene glycol, are used in latex paints (i.e., as opposed to
Emissions of TMPD-MIB from Latex Paint 1055

aliphatic and aromatic hydrocarbons used in oil-based paints). The glycols


added to latex paint also provide for freeze protection. Most paints also con-
tain one or more additives that serve special purposes. These additives can
include biocides, coalescing agents, defoaming agents, thickeners, and sta-
bilizers, among others. A key additive in most latex paints, and the focus of
this 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 facilitates
complete fusion when paint dries. It also enhances scrub resistance, color
development, and packaging stability.

Volatile Components of Latex Paint


The percentage by weight of the non-volatile components (resin + pigment)
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

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., Chang
et al., 1997; Fortmann et al., 1993; Sparks et al., 1999a). Water dominates the
overall 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 classified
as both volatile and semi-volatile organic compounds.
The actual composition of TVOC can vary significantly between latex
paints (e.g., semi-gloss latex paints are typically comprised of greater TMPD-
MIB weight fractions than are flat latex paints). Censullo et al. (1996) reported
TVOC composition analysis for 34 latex paints. They observed a wide range
of (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, or
BEE), ethylene glycol (EG), propylene glycol (PG), and TMPD-MIB. Based
on their analysis, Censullo et al. (1996) proposed the following compos-
ite group species profile for water-based paints as percentage by weight of
TVOC: TMPD-MIB (35%), PG (33%), EG (17%), BEE (6%), other (residual).
However, this was intended as a gross species profile that could be used for
outdoor emissions inventories. It does not account for the significant vari-
ability between waterborne paints. For example, the percentage by weight
TMPD-MIB contribution to TVOC varied from 0% to greater than 70%.
Several other researchers have reported TMPD-MIB of between 20 and
33% of TVOC on a percentage by weight basis for latex paints intended for
interior 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 close
to, the central tendency reported by Censullo et al. (1996). In each of the
aforementioned studies, the dominant (S)VOCs were consistent with those
reported by Censullo et al. (1996). Ethylene glycol and TMPD-MIB, in that
order, dominated TVOC composition, with lesser amounts of BEE and PG
1056 R. L. Corsi and C.-C. Lin

in each case. The reported contribution of EG was much greater, and the
reported contribution of PG much lower, than that reported by Censullo et
al. (1996).
Assuming a TMPD-MIB content of 20–33% by weight of TVOC and a
TVOC content of 2–5% by weight yields an estimated TMPD-MIB content
of 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 primer
sealer, flat latex paint, and semi-gloss latex paint used in large chamber
experiments. The percentage by weight compositions of TMPD-MIB for these
three products was 1.5%, 0.9%, and 0.7%, respectively. The percentage by
weight compositions of EG were 3.6%, 2.0%, and <0.7%, respectively. The
percentage by weight compositions of PG were <0.8%, <0.8%, and 3.1%,
respectively. Fortmann et al. (1993) reported the mean composition of three
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

replicates 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 four
replicates 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 the
basis for experiments/model development by the USEPA, which will be
described 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 of
later discussion, the properties of several other components of latex paint

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

Aqueous Pvp
solubility [mm
Molecular CAS MW [mg/L] Hg] Tb Tf log10
Compound formula # [g/mol] @ 20◦ C @ 20◦ C [◦ C] [◦ C] (Kow )

BEE C8 H18 O3 112-34-5 162.2 Miscible 0.02 231 −68 0.15 to 0.40
EG C2 H6 O2 107-21-1 62.1 Miscible∗ 0.05 198 −12.7 −1.93
PG C3 H8 O2 57-55-6 76.1 Miscible∗ 0.2 188.2 −59 −1.40 to −0.30
TMPD- C12 H24 O3 25265-77-4 216.4 858 0.01 254 −50 3.47
MIB (18–22◦ C)
Water H2 O 7732-18-5 18 — 17.54 100 0 —
For each chemical, the relevant temperature was not reported for Kow . Values for water are widely
tabulated and published extensively.
All values except for TMPD-MIB and water are from Verschueren (1996) and ∗ Agency for Toxic Substances
and 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/melting
point, Kow = octanol/water partition coefficient.
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 the
isomers (x = 1), the ester bond involves the hydroxy group of the C1 atom
in the diol moiety, and in the second isomer (x = 3), it involves the hydroxy
group 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 (≈112
ppb 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 partition
coefficient. The former indicates a relatively slow rate of evaporation. The
latter 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.
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

Emissions of TMPD-MIB from Latex Paints: Conceptual Development


It is worthwhile to consider the major processes that may affect TMPD-MIB
emissions, even in qualitative terms, as a prelude to the interpretation of
published research results described later in this paper.
The release of TMPD-MIB from latex paint to indoor air is a dynamic
process that is influenced by the continuously changing nature of the paint
following application to a material (i.e., from initial wet paint composition
to ultimate dry paint film). However, the emissions process is generally
described by two major stages. The first is a relatively short stage, during
which the paint is treated as a wet film. The second is a prolonged stage
in which the paint is treated as a dry film. These two stages are depicted in
Figure 1. In the wet film, TMPD-MIB can exist in three major components:
dissolved in the aqueous phase, as a separate pure phase, and adsorbed
to (or absorbed by) resin particles. It may also volatilize to room air, or
partition into the gas phase with subsequent diffusion into pores of the
substrate to which the paint is applied. The presence of co-solutes such as
ethylene glycol (EG), propylene glycol (PG), and 2-(2-butoxyethoxy)ethanol
(BEE) may have some influence on the solubility of TMPD-MIB in the liquid
portion of the paint emulsion. However, the solubility of TMPD-MIB is likely
to remain very low. As described previously, the percentage by weight of
TMPD-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 range
would translate to 8,000 to 34,000 mg/L if all of the TMPD-MIB could be
dissolved 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 small
fraction of the TMPD-MIB mass in paint is dissolved in the liquid phase.
1058 R. L. Corsi and C.-C. Lin

FIGURE 1. Schematic depicting latex paint applied to a building material (substrate), includ-
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

ing emissions from (a) a wet film and (b) ultimate dry film absent of water. The lines internal
to the substrate depict VOC fate mechanisms internal to the substrate, including molecular
diffusion, 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 softening
of resin particles, it is reasonable that a majority of its mass exists adsorbed
to, or absorbed within, the large mass of organic resin present in the paint.
Nevertheless, the fraction dissolved in the aqueous phase and any separate
pure phase will play some role in the emissions process during the drying
(wet film) stage. The mechanistic behavior of corresponding emissions is
complicated by the fact that water evaporates faster than the other solvents
or TMPD-MIB, thus leading to changes in co-solvent effects and relative
partitioning of TMPD-MIB between the liquid and particle phases of the
emulsion.
The dry film stage (see Figure 1b) will exist following complete evapo-
ration of water. The dry film is comprised of resin-pigment, residual organic
solvents, and TMPD-MIB. The process by which TMPD-MIB emissions occur
from the dry film are not well understood, but have been reported to persist
long after paint drying (Clausen et al., 1991; Hodgson et al., 2000; Lin &
Corsi, 2007; USEPA, 2002), and a majority of emissions have been noted to
occur from the dry film (USEPA, 2002). It is likely that molecular diffusion
through the dried paint film, and possibly the substrate (Lin & Corsi, 2007),
plays a major role in prolonged emissions from paint after it reaches the
dry film stage. Clausen et al. (1991) hypothesized that either diffusion or
evaporation can dominate emissions after the film forms, depending on the
specific 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
Emissions of TMPD-MIB from Latex Paint 1059

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

SUMMARY OF PAST RESEARCH

Empirical observations of TMPD-MIB emissions from latex paint have been


reported in the published literature. Much of this is related to the use of small
continuous flow laboratory chambers constructed of stainless steel or glass
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

(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 et
al., 1999a; Van der wal et al., 1997) or the Field and Laboratory Emission
Cell (FLEC) (De Bortoli et al., 1999; Roache et al., 1996; Wolkoff, 1998)
to quantify area-specific emission rates. The relevance of many laboratory
experiments to real-world conditions has been limited due to the extensive
use 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 number
of 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 sorptive
sinks (i.e., other materials to which TMPD-MIB can adsorb to and desorb
from).
Every study reviewed for this paper, whether conducted in the labora-
tory or field, involved the collection of TMPD-MIB on a solid adsorbent with
subsequent thermal or chemical extraction and analysis via either GC/FID or

R
GC/MS. Tenax -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; Sparks
et 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 standard
charcoal tubes when extracted by CS2 . Fortmann et al. (1993) used XAD
resin extracted with methylene chloride, but noted the need for more per-
formance data related to different sorbents. De Bortoli et al. (1999) reported
1060 R. L. Corsi and C.-C. Lin

very good inter-laboratory results for TMPD-MIB emissions associated with


an 18 laboratory/10 country European study involving three materials, one
of which was a latex paint. However, they concluded that analytical errors
were by far the greatest cause for variance in results between laboratories.
Past research has provided some insight into the order of magnitude
of TMPD-MIB emissions following latex paint applications to both idealized
and 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 been
reported 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 on
the effects of the method of application and subsequent film thickness on
TMPD-MIB emissions (Clausen, 1993; Clausen et al., 1991; Fortmann et al.,
1993; Yu & Crump, 1999). Several research teams have reported on the
effects of environmental conditions on TMPD-MIB emissions, with variations
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

in air temperature, relative humidity, and air speed above the wet and/or
dry 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). Relevant
findings 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 either
indoor or outdoor air, in the published literature. Girman et al. (1999) re-
ported VOC concentrations in 56 public and private office buildings as part
of 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 of
TMPD-MIB concentrations was 0.5 to 28 µg/m3 (∼0.1–3 ppb). Hodgson
et al. (2000) completed a field study involving 11 homes (four manufac-
tured homes and seven site-built homes). The four manufactured homes
were 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 ppb
and 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 as
follows: 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 after
painting the walls of a bedroom with latex paint. All of the walls were first
covered with new gypsum board. Latex paint was then applied with a roller
to an area-average mass coverage of 155 g/m2 . The average air exchange rate
of the test house was 0.37/hr, with a temperature of 22.2◦ C and a relative
humidity of 70%. The approximate (read from log axis) concentration of
Emissions of TMPD-MIB from Latex Paint 1061

TMPD-MIB at 200 and 1400 hours (after paint application) in the painted
bedroom was 10 ppb and 3 ppb, respectively. Similar concentrations were
observed in a hallway near the bedroom at the same times. The value of 3
ppb nearly two months after paint application was generally consistent with
the ranges reported by Hodgson et al. (2000) for residential buildings.
Shields and Weschler (1992) completed a three-year study of VOCs in
a telephone switching station in Wisconsin, USA. TMPD-MIB was observed
in room air during and following two renovation events in the building. The
measured concentrations ranged from highs of approximately 1 to 2 ppb
during renovation to between 0.6 and 1.1 ppb three months later. These
relatively low values were likely due to high ventilation rates in the build-
ing during and following renovation events. During the first and second
renovation events, supply air was comprised of 37% and 68% outside air,
respectively.
Norback et al. (1995) completed an exposure analysis of 20 house
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

painters 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 be
1,680 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 ppb
at 20◦ C and 1 atmosphere) (GSD = 7.0 µg/m3 ). Thus, while some painters
were exposed to relatively high concentrations of TMPD-MIB, the geomet-
ric mean exposure concentration was on the same order of magnitude as
background concentrations reported by Girman et al. (1999) for commer-
cial office buildings and Hodgson et al. (2000) and Sparks et al. (1999b) for
residential homes several months after painting.
Kim et al. (2007) measured TMPD-MIB concentrations in 23 classrooms
of eight schools in Sweden. The mean concentration was 0.89 µg/m3 , with
a range of 0.07 to 4.41 µg/m3 . The highest concentrations were measured in
newer 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 the
time of application than those studies described above. A latex primer sealer
was applied to gypsum board and plywood. Flat and semi-gloss latex paints
were then applied to gypsum board and plywood, respectively. The base
condition had an air exchange rate of 2/hour for several hours, followed
by an extended period of air exchange at 0.5/hr. Chamber air exchange
rates and air mixing intensity were varied, but had little effect on measured
concentrations 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 to
179 ppb at 240 hours, levels much greater than those reported by Norback
1062 R. L. Corsi and C.-C. Lin

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


applications.

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 paint
composition, 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 on
flow-through small chamber experiments and were calculated based on an
assumption of pseudo-steady-state conditions at the time of sampling (48 and
96 hours after product application). The highest emission factors were for a
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

latex primer sealer and flat latex paint applied to gypsum board. Emission
factors for paints applied to gypsum board decreased significantly from hour
49 to 96. This was not the case for semi-gloss paint applied to plywood.
Emission factors gleaned from six other published studies are reported
in Table 3. Four of these involved applications of latex paints to gypsum
board, 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 chamber
experiments.

Coverage Emission factor Emission factor


of 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 6.94


LPS2 Gypsum board 110 2.64 1.97 24.0 18.0
FLP1a Gypsum board 99 2.02 0.26 20.8 2.59
FLP1b Gypsum board 104 2.32 0.30 22.4 2.93
FLP2 Gypsum board 111 1.43 0.28 12.9 2.48
FLP3 Gypsum board 163 2.86 1.42 23.8 8.73
SGLP1 Plywood 164 0.34 0.26 2.07 1.56
SGLP2 Plywood 129 0.88 0.82 6.88 6.33
SGLP3a Plywood 169 0.93 0.89 5.55 5.30
SGLP3b Plywood 160 0.72 0.54 4.39 3.33
LPS2 + FLP3 RH = 30% Gypsum board 187 5.61 3.83 30.7 20.9
LPS2 + FLP3 RH = 50% Gypsum board 156 6.26 3.20 40.0 20.5
LPS2 + FLP3 RH = 70% Gypsum board 163 5.32 2.68 32.7 16.5
LPS2 + 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.
Emissions of TMPD-MIB from Latex Paint 1063

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

Emission factor Time


Material mg/m2 -hr (units noted) Note Reference

Gypsum board 0.002 11 months Estimated based on Chang et al.


read off log (1997)
concentration plot
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 4.8 48 hours 18 lab study: CV = De Bortoli et al.
impermeable 21–25%, small (1999)
substrates chambers and
FLEC used
Glass 29 3 hours Yu and Crump
(1999)
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

3.3 27 hours
1.4 76 hours

Nevertheless, it appears that significant reductions in area-specific emission


rates for TMPD-MIB occur over time. The last two entries in Table 3 involve
applications of latex paints to impermeable substrates such as stainless steel,
polyester sheets, and glass. The time scales are similar to those reported by
Hodgson (1999) for gypsum board. Interestingly, the area-specific emission
rates for the impermeable materials for the first 27 to 48 hours are reasonably
similar to those reported by Hodgson (1999) for gypsum board, although a
direct 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 test
chamber 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 small
flow-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 were
observed for up to 15 months.
Chang et al. (1997) applied latex paint to gypsum board and measured
concentrations in a small flow-through chamber over an 11-month period.
The concentration of TMPD-MIB was greater than that of ethylene glycol for
the 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 in
off-gas after one year.
1064 R. L. Corsi and C.-C. Lin

Sparks et al. (1999b) applied latex paint to gypsum board in a USEPA


test house. TMPD-MIB concentrations in house air decreased significantly
over the first 400 hours. However, they were relatively constant from 400 to
4000 hours after application, suggesting either relatively constant prolonged
emissions from the dry paint and/or constant re-emission rates from interior
sinks 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 is
typically defined as the cumulative mass emitted to air divided by the mass of
the compound initially applied to the substrate. The recovery of total volatile
organic compound (TVOC) mass two weeks after latex paint is applied to
stainless steel has been observed to be nearly 100% (Sparks et al., 1999a, and
references provided therein). Similarly, recovery of TMPD-MIB two weeks
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

after latex paint is applied to stainless steel has been observed to be on the
order of 90% (Chang et al., 1997; Guo et al., 1996).
Only a few TMPD-MIB recoveries have been reported for latex paint
applications to gypsum board; no published reports of TMPD-MIB recovery
following latex paint applications to wood, concrete, or stucco were identi-
fied during this review. A summary of published results for painted gypsum
board is presented in Table 4. Note that the recovery period never exceeded
two weeks in these studies except for Lin and Corsi (2007) and, as such,
long-term recoveries of TMPD-MIB from actual building materials are still
deficient. Over the three two-week studies, the range of TMPD-MIB recov-
eries from gypsum board is relatively large (27 to 60%), with differences
being difficult to ascertain based on available data. All three two-week pe-
riod recoveries were based on latex paint applied to gypsum board placed
in small stainless-steel flow-through chambers with similar air exchange rate,
relative humidity, and temperature. The compositions of the paints used by
Chang et al. (1997) and Guo et al. (1996) were similar in terms of TVOC and
percentage by weight contribution of TMPD-MIB and other (S)VOCs. The
composition of paint used by Roache et al. (1996) was not described in the
source paper.

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

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)
Emissions of TMPD-MIB from Latex Paint 1065

Lin and Corsi (2007) observed TMPD-MIB emission rates for periods
as long as 15 months. The airborne recoveries of TMPD-MIB were a strong
function of the type of paint and substrate. Recoveries in air of approximately
50% (low-pigment volume paint) to 90% (high-pigment volume paint) were
observed after 15 months for regular applications to gypsum board. Lin and
Corsi (2007) were the first to complete solvent extractions of the painted
substrate and to perform a mass closure analysis of TMPD-MIB. Results
for 10 gypsum board specimens yielded a 96 ± 6% mass closure (i.e., on
average, nearly all of the TMPD-MIB not emitted to air was typically observed
in the substrate extracts). A fairly high percentage (≈40%) of the TMPD-
MIB applied to gypsum board in low pigment volume latex paint (tending
toward semi-gloss paint) was sorbed within the gypsum board matrix, even
15 months after paint applications. Low-level emissions to air after three
months following paint applications were dominated by releases from the
dry paint film, and not in the form of TMPD-MIB diffusing from the interior
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

of the gypsum board.


Based on a review of existing literature, there is a clear need for more
long-term TMPD-MIB mass recovery data following applications to products
other 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 on
impermeable (non-porous) materials such as stainless steel, glass, and alu-
minum (Clausen, 1993; Clausen et al., 1991; Fortmann et al., 1993). Even
more recently, several researchers have continued to use substrates such as
glass (Yu and Crump, 1999), aluminum (Van der wal et al., 1997), polyester
sheets, stainless steel, and aluminum (De Bortoli et al., 1999). In fairness,
the intent of these studies was not necessarily to provide realistic emissions
data for indoor paint applications; for example, the study by De Bortoli
et al. (1999) was intended as an inter-laboratory comparison of methods
used to characterize emissions from building materials.
More recently, there has been greater emphasis on the use of actual
indoor substrates to study emissions from latex paint. Most of the published
accounts of latex paints applied to actual building materials have focused
on 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 to
determine emission rates. For this review, only one set of results for latex
paint applications to a wood product (plywood) was found (Hodgson, 1999).
The substrate to which paint is applied can have a significant influence
on 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
1066 R. L. Corsi and C.-C. Lin

substrates, including glass, gypsum board, and gypsum board covered with
wood chip wallpaper. Substrate effects were relatively small for the non-polar
n-alkanes and aromatics. Although TMPD-MIB was not included in the paint
mixture, 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 to
each of the two gypsum board substrates relative to emissions from glass.
Krebs et al. (1995) also emphasized that substrates can have a significant
influence 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 (Chang
et 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 substrates
have been observed. In general, the peak concentration in chamber air oc-
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

curs sooner and is smaller when paint is applied to gypsum board; also,
as noted previously, mass recovery of TVOC or individual components are
lower 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 the
substrate 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 paint
components, the emissions retardation cannot simply be due to shielding
of mass from evaporative processes by liquid paint “soak” into the pores of
a substrate such as gypsum board. Instead, emissions may be reduced by
sorptive interactions between polar compounds and gypsum board while
in either the liquid phase (liquid-solid interactions) or following liquid-gas
partitioning and gas-phase diffusion/adsorption processes within the pores
of the substrate.
Importantly, it appears that the porosity and/or polarity of a substrate is
an important factor in determining oxygenated (S)VOC emissions following
paint application. Applications of the same latex paint to different non-
porous materials (polyester sheets, glass, aluminum) had little impact on
area-specific emission rates during a large inter-laboratory comparison (De
Bortoli 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 gypsum
board, gypsum board that had been previously painted, and eight-year-old,
painted gypsum board. Emissions of TVOC following fresh paint applications
were 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
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; experiments
were 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 for
these compounds. These results are significant and warrant further research
given extensive applications of paint to concrete in some parts of the world.
For example, in China, a majority of architectural paint products, including
latex paints, is applied to concrete (Avery, 2004).
It is clear that the choice of substrate has a significant influence on
emissions from latex paint, and that the emissions process is retarded for
TMPD-MIB and other polar components when latex paint is applied to gyp-
sum board. While more research is needed to fully understand the nature of
the emissions process following paint applications to gypsum board, wood
products and concrete are also deserving of additional attention. Only one
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

study 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 test
substrates for emissions testing (USEPA, 2002). Most published studies have
involved the application of latex paint to gypsum board with rollers (Chang
et 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) used
brushes to apply latex paint to glass.
The most important factor associated with paint application appears to
be the thickness of the paint film. Clausen et al. (1991) found that long-term
emissions are reduced by reducing the paint film thickness on tin-plated
steel, and that normalization of the emissions by film thickness is critical
for avoiding incorrect conclusions when comparing emissions from paint.
For TMPD-MIB, peak chamber concentrations increased with increasing film
thickness (22, 28, and 56 mm) on tin-plated steel, and decayed at a much
slower rate for the specimen with the thickest film (Clausen, 1993).
Nominal initial wet-film thicknesses for paints applied to gypsum board
using 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 Clausen
et al. (1991) reported final dry film thicknesses of between 22 and 63 µm
using a Twintector Elcometer device for latex paint applied to tin-plated
steel.
In the future, it might be possible to gain additional insights into the
emissions process by varying the thickness of paint applied to gypsum board
and measuring temporal differences in both emissions and mass recovery of
individual components. For example, one would not expect significant dif-
ferences in emissions from the wet film if the film composition is relatively
1068 R. L. Corsi and C.-C. Lin

homogeneous unless a greater fraction of film mass is adsorbed by the


substrate for thinner films. Similarly, a lower long-term recovery from appli-
cations involving thick films might suggest the importance of the dry-film
as a barrier for diffusion of TMPD-MIB and other components out of the
substrate 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 may
affect TMPD-MIB emissions. For example, for an otherwise equivalent paint
composition, an increase in TMPD-MIB weight fraction should lead to an
increase in TMPD-MIB emissions, presumably during both paint drying and
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

thereafter.
The effects of variations in other paint components, particularly the
solvents ethylene glycol and propylene glycol, may also affect TMPD-MIB
emissions. As described elsewhere in this paper, increases in EG and PG
weight fractions may increase the solubility of TMPD-MIB in the liquid phase,
thereby increasing the TMPD-MIB mass available for volatilization during
paint drying. Increasing values of EG and PG content may also influence
the nature and extent of TMPD-MIB adsorption within a substrate through
multi-component competition for sorption sites. However, the importance
of these effects cannot be gleaned from previous publications related to
emissions from paint products, and is an area in need of additional research.
After paint has dried, long-term emissions of TMPD-MIB are presumably
dependent, at least in part, on the nature of TMPD-MIB molecular diffusion
through and/or out of the dried paint film. The nature of a dried paint film
on the effective diffusion coefficient of TMPD-MIB may vary depending on
the solids composition of a paint mixture. For example, flat latex paints
tend to have greater amounts of “filler” material relative to semi-gloss latex
paints that often contain a much larger fraction of acrylic resins. Thus, the
effective diffusion coefficient through the dry film of flat latex paint may be
considerably different than that through semi-gloss latex paint. Additional
research is needed to assess the significance of differences in dry paint film
properties 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 that
have been studied are air temperature, relative humidity (RH), and air speed
above the painted surface.
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 Laboratory
R
Emission Cell (FLEC) to measure the concentration-time profiles of 1,2-
propanediol and TMPD-MIB over 250 days. The effects of temperatures of
23, 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. For
1,2-propanediol at 60◦ C, the concentration-time profile dropped below that
observed 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-specific
emission 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 with
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

increasing temperature at 30% and 50% RH. However, a significant increase


in emissions of TMPD-MIB was observed with increasing temperature at the
elevated relative humidity (70%). A four-fold increase in emission rate (∼51
to 216 µg/m2 -hr) was observed with a 10◦ C increase in temperature (18 to
28◦ C) at 70% RH.
Van der wal et al. (1997) applied a latex paint to tin-plated steel and
measured concentrations of TMPD-MIB over a two-week period in two small
laboratory chambers. One chamber was operated at 23◦ C and the other at
30◦ C. The relative humidity was maintained at 45% in each chamber. The
peak concentration of TMPD-MIB was approximately twice as great at the
elevated 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% to
50% RH had a significant impact on emissions of 1,2-propanediol (higher
emissions at higher RH), there was no observable effect on emissions of
TMPD-MIB over 25 days. This result is consistent with findings by Hodg-
son (1999), who observed little effects of changes in RH on emissions of
TMPD-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 as
relative 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 have
a 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
1070 R. L. Corsi and C.-C. Lin

consistent with those measured by Sparks et al. (1999b) near the walls of a
USEPA test house. He observed little differences in emissions of TMPD-MIB
over this range of air speeds. This result is consistent with Hodgson and
Shimer (1999), who observed little effect of increasing air exchange rate or
mixing intensity of air (additional fans) on the concentration of TMPD-MIB
during large chamber experiments.

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


Sorptive interactions with chamber or indoor materials after being emitted
from paint may affect determination of emission rates of TMPD-MIB, as well
as its persistence within building environments. For example, Clausen et al.
(1991) noted potential problems with TMPD-MIB adsorption and prolonged
desorption from the walls of small chambers.
Chang et al. (1998) exposed 17-year-old painted gypsum board removed
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

from a USEPA test house to relatively high gaseous concentrations of TMPD-


MIB, BEE, EG, and PG, followed by a 300-hour purge/desorption stage. They
determined a mass recovery of only 18% for TMPD-MIB. Recoveries for BEE,
EG, and PG were somewhat lower, a trend consistent with other studies
involving 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 carpet
exposed to TMPD-MIB. The rate constants were determined based on the
classic linear surface sorption model:

Rsin k = ka C − kd Msin 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 of
the chemical of interest adsorbed to the sink material (mg/m2 ). The equilib-
rium partition coefficient (Keq ) for each material was determined as the ratio
of 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 relative
to EG, PG, and BEE for carpet. The resulting Keq is an approximate order
of magnitude lower for TMPD-MIB than for the other components of latex
paint. This suggests that the strong polar-polar interactions between gyp-
sum board and each polar compound effectively dampens any differences
in sorption parameters between the compounds, while the same is not true
for the less polar surfaces of carpet components. Interestingly, the results
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 ± 0.083


kd (1/hr) 0.0048 ± 0.00075 0.016 ± 0.0023
Keq (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 one
standard deviation.

of Won et al. (2000) suggest that because TMPD-MIB has a lower vapor
pressure than the other constituents of latex paint, its equilibrium partition
coefficient should be greater than those of the other compounds for carpet.
The opposite was noted by Sparks et al. (1999a).
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

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 indoor
environments. 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, Shields
and Weschler (1992) measured the concentrations of each TMPD-MIB isomer
over the course of three months after latex paint applications in a telephone
switching station. After each of two painting events, they observed a greater
relative 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 that
indicate measurable first-order hydrolysis rates for TMPD-MIB in alkaline so-
lutions (Roser, 1992). However, additional research is needed, particularly
over 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 the
ultimate 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 for
the emissions process have yet to be developed, and data to support a
highly mechanistic model of the emissions process do not exist in the pub-
lished literature. Nevertheless, several researchers have developed empirical
and semi-empirical models to estimate species emissions following paint
1072 R. L. Corsi and C.-C. Lin

applications, and have reported model parameters for TMPD-MIB. Those


models are described in this section, along with several salient features asso-
ciated with their development and parameters relevant to TMPD-MIB, where
appropriate. The reader is referred to the original publications for details
related to model derivations and experimental methodologies used to deter-
mine model parameters.
A simple first-order decay equation has been used to estimate VOC
emissions 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 variables
are as defined above.) The empirical nature of this model precludes its ef-
fective application outside of the conditions for which the decay constant
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

is determined experimentally. A significant problem with Eq. (2) is that if


the decay constant is based on a limited amount of short-term data, the
model 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-controlled
phase).
To overcome problems associated with the prediction of long-term emis-
sions, some researchers have employed a double-exponential decay model
to estimate VOC emissions from architectural coatings (e.g., Chang et al.,
1997; Wilkes et al., 1996):

Ea (t) = E1a (0)e−k1 t + E2a (0)e−k2 t (3)

where E1a (0) and E2a (0) = the initial (t = 0) area-specific emission factors
for phases 1 and 2, respectively (mg/m2 -hr), k1 and k2 = the emission decay
rate constants for phases 1 and 2, respectively (1/hr), and all other variables
are as defined above. In this model, short-term emissions are meant to be
accounted 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 the
right-hand-side of the equation. As such, k1  k2 .
While Eq. (3) can provide better estimates than Eq. (2) for long-term
emissions, it still suffers from the limitations described above in terms of
its empirical nature (i.e., it cannot effectively be used to estimate emissions
for conditions other than those used to experimentally determine k1 and
k2 ). Wilkes et al. (1996) suggested the possibility of overcoming part of this
limitation by correlating k1 to chemical vapor pressure and k2 to molecu-
lar weight of the emitted component. However, this approach still fails to
account for the changing composition of a wet paint film (changing mole
fractions of emitting constituents), the nature of the dried paint film (which
varies in time and between paints), and environmental conditions.
Emissions of TMPD-MIB from Latex Paint 1073

Clausen et al. (1991) and Clausen (1993) presented Eqs. (4) and (5) to
predict emissions following latex paint applications:
  k t
kε1 − ε1L
Ea (t) = M(0) e (4)
L
   
kD1 − kLD12 t
Ea (t) = M(0) e (5)
L2

where kε1 = the rate constant for evaporation controlled emissions for
L = 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). (All
other 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 film
itself). It is based on application of Fick’s first law with an assumed linear
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

concentration gradient through a laminar boundary layer above the paint.


The rate constant for evaporation-controlled emissions includes the diffusion
coefficient in air for the chemical of interest as well as the thickness of the
laminar 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’s
second law, with an assumed zero concentration for the chemical of interest
at the film surface. It is further assumed that there is a zero flux (no flux
condition) at the interface of the paint film and underlying substrate. While
this assumption might be valid for impermeable surfaces such as aluminum
and glass, it is likely not valid for porous materials such as gypsum board or
concrete.
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 and
kD1 /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 both
short-term and long-term chemical emissions:

αfD MD (0)e−2fD t
1/2
−kt
Ea (t) = MV (0)ke + (6)
t1/2
where MV (0) = initial (t = 0) mass of chemical available for emissions by
evaporation (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 defined
previously. The term α is an adjusting factor that is needed to account for
the fact that diffusion-controlled emissions cannot become important until
the paint film is dried. Guo et al. (1996) selected the following equation
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 differential
equation to represent the time and one-dimensional space-distribution of a
chemical diffusing through a layer of thickness λ, as described by Treybal
(1968):

0.632D1/2
fD = (8)
λ

where D = the effective diffusion coefficient through some medium (m2 /hr)
and λ = the thickness of the diffusion later (m).
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

Sparks et al. (1999a, 1999b) developed and applied what is arguably


the most mechanistic of models published for estimation of chemical emis-
sions following paint applications. The model is effectively a coupling of the
following three equations:
   
MV (t) MV (t) 2 fD MD (t)
Ea (t) = km CVp − C(t) + 1 − (9)
MV (0) MV (0) t1/2
 
dMV (t) MV (t)
= −km CVp − C(t) (10)
dt MV (0)
 
dMD (t) MV (t) 2 fD MD (t)
=− 1− (11)
dt MV (0) t1/2

where km = gas-phase mass transfer coefficient (m/hr), CVp = vapor pressure


for 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 for
emissions by molecular diffusion (mg/m2 ). (All other variables are as defined
above.) The gas-phase mass transfer coefficient (km ) can be estimated based
on 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 in
Table 6. The small number of studies for which TMPD-MIB parameters are
available for a common substrate, and the different conditions for which
model parameters were derived, precludes a detailed comparison of model
predictions.
Emissions of TMPD-MIB from Latex Paint 1075

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

Paint/substrate Experimental
Equations properties conditions Model parameters Reference

2 6.3 L chamber M(0) (mg/m2 )


Al plates. Acrylic paint: Van der wal
composition not (0.95 ACH) 23◦ C/30◦ C et al.
provided. 24-hour ·Isomer 1 = 64/41 (1997)
drying ·Isomer 2 = 101/66
period prior k (hr−1 × 10−3 )
to measure- ·Isomer 1 = 59/230
ments. ·Isomer 2 = 41/260
T = 23◦ C &
30◦ C; 45%
RH
3 Stainless steel and 53 L chamber Stainless steel Chang et al.
gypsum board. Paint: (0.5 ACH) (short-term) (1997)
Nominal wet film = 23◦ C; 50% RH ·E1a (0) = 30 mg/m2 -hr
100 µm. ·E2a (0) = 0 mg/m2 -hr
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

TVOC = 45.4 mg/g; ·k1 = 0.0169/hr


PG = 2.32; EG = 24.0; Gypsum board
BEE = 4.98; ·E1a (0) = 29.7
TMPD-MIB = 13.5; mg/m2 -hr
diethylene glycol = ·E2a (0) = 15.9
0.59 mg/m2 -hr
·k1 = 0.795/hr
·k2 = 0.0317/hr
3 Pre-painted gypsum Small k1 = 0.944/hr Wilkes et al.
board. chambers k2 = 0.013/hr (1996)
Latex paint: EG (54%), E1a (0) = 1.46 mg/hr
PG, BEE, TMPD-MIB E2a (0) = 0.22 mg/hr
(30%)
2, 4, 5 Tin-plated stainless 234 L chamber Values for L = 22/28/56 Clausen
steel. (0.25 ACH) µm (1993)
Waterborne paint = 23 ± 0.6◦ C; 45 ·M(0) (mg/m2 ): 35/ 46/
white spirit; ± 3% RH 410
1,2-propanediol, BEE; ·k (hr−1 ): 0.015/ 0.016/
TMPD-MIB (fractions 0.0048
not provided). Dry ·kε1 (µm/hr): 0.33/ 0.46/
film thickness = 22, 0.27
28, 56 µm. ·k D1 (µm2 /hr): 7/ 13 /15
6, 7 Gypsum board. Paint: 53 L chamber k = 0.0635 /hr Guo et al.
4.5% TVOC: EG = (0.5 ACH) MV (0) = 404 mg/m2 (1996)
53%; TMPD-MIB = 23◦ C; 50% RH Md (0) = 1,465 mg/m2
30%; BEE = 11%; fd = 0.00173 /hr1/2
PG = 5%; diethylene
glycol = 1%
9–11 Gypsum board. Paint: 53-L chamber CVO = 20 g/m3 Sparks et al.
TVOC = 45 mg/g; (0.5 ACH) MV (0) = 496 mg/m2 (1999a)
EG = 24; TMPD 23◦ C; 50% RH fd = 0.0012/hr1/2
-MIB = 13; BEE = 5;
PG = 2; diethylene
glycol = 2
1076 R. L. Corsi and C.-C. Lin

SUMMARY AND RESEARCH NEEDS

The existing literature related to TMPD-MIB emissions from latex paint is


sparse and generally insufficient to make highly certain conclusions regard-
ing the nature of the emissions process or the ultimate fate of TMPD-MIB. The
same is true for other volatile components of latex paint, and the small base
of published data related to TMPD-MIB precludes any attempts to extrapolate
information to the other compounds. Nevertheless, the following observa-
tions, conclusions, and recommendations for future research are made based
on this review:

r Published measurements of TMPD-MIB emissions following latex paint


applications are generally limited by a lack of analysis of initial paint com-
position, as well as numerous reports of paint applications to impermeable
substrates such as stainless steel and glass.
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

r Long-term emissions data for TMPD-MIB are sparse, but existing data do
suggest continued emissions for up to at least a year after paint application.
r The range of TMPD-MIB recoveries in air following paint applications
to gypsum board is relatively large, with differences being difficult to
ascertain based on available data.
r The substrate to which paint is applied can have a significant influence
on (S)VOC emissions, including TMPD-MIB. This is particularly true when
comparing emissions from impermeable materials (e.g., stainless steel)
with emissions from porous materials such as gypsum board or concrete.
r It is clear that either liquid-solid and/or gas-solid sorptive interactions
between TMPD-MIB and gypsum board retard short-term emissions and
prolong the emissions process relative to emissions from impermeable
substrates.
r The most important factor associated with paint application appears to be
the thickness of the paint film. However, additional work is needed to
assess the effects of paint composition on TMPD-MIB emissions.
r The effects of temperature and relative humidity on TMPD-MIB emissions
from latex paint are not well understood and may be coupled. It does not
appear that air speed above painted surfaces has a significant influence
on TMPD-MIB emissions.
r The migration and storage of TMPD-MIB into gypsum board is observed
to be significant; absolute mass storage in the substrate is typically greater
than that in the dry paint film after six months or more.

There is a clear need for long-term mass recovery data associated with
TMPD-MIB following applications to wood products, concrete, and other
cementatious materials.
Significantly more research, as well as new research approaches, is
needed to improve the existing knowledge base associated with TMPD-MIB
Emissions of TMPD-MIB from Latex Paint 1077

emissions from latex paint and to ultimately allow for the development of
improved 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. Special
thanks are given to Dr. Bruce Gustafson of Eastman Chemical and Dr. David
Morgott, as well as to Mr. Robert Avery for their expeditious provision of in-
formation when requested. During this work, Chi Chi Lin was supported by
Grant No. T42CCT610417 from the Southwest Center for Occupational and
Environment Health, as supported by the National Institute for Occupational
and Environmental Health (NIOSH)/Centers for Disease Control and Preven-
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

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, contract
no. 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 to
the California Air Resources Board, contract no. 93–319.
Chang, J.C.S., Sparks, L.E., Guo, Z., and Fortmann, R. (1998). Evaluation of sink
effects on VOCs from latex paint. Journal of the Air & Waste Management
Association, 48, 953–958.
Chang, J.C.S., Tichenor, B.A., Guo, Z., and Krebs, K.A. (1997). Substrate effects on
VOC emissions from a latex paint. Indoor Air, 7, 241–247.
Clausen, P. (1993). Emission of volatile and semi-volatile organic compounds
from 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 of
volatile 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.
1078 R. L. Corsi and C.-C. Lin


R
Eastman Chemical Company. (2003). Material safety data sheet: Texanol ester
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 humidity
on chemical sensory emissions from building materials. Indoor Air, 9, 193–201.
Fortmann, R.C., Sheldon, L.S., Keever, J.T., Whitakerand, D.A. (1993). Comparison
of 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 wall
paints—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. Proceedings
of 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 to
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

gypsum 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 National
Laboratory, 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-built
houses. Indoor Air, 10, 178–192.
Hodgson, A.T., and Shimer, D.A. (1999). Techniques for reducing exposures to
volatile 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 in
schools—associations with asthma and respiratory symptoms in pupils. Indoor
Air, 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 and
chemical 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. Engineering
Solutions to Indoor Air Quality Problems, a specialty conference sponsored by
the Air & Waste Management Association and the United States Environmental
Protection 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.
Emissions of TMPD-MIB from Latex Paint 1079


R
Lin, C.C., and Corsi, R.L. (2007). Texanol 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 standards
or guidelines for chemicals. Indoor Air, 7, 17–32.
Norback, D., Wieslander, G., and Edling, C. (1995). Occupational exposure to volatile
organic compounds (VOCs), and other air pollutants from the indoor application
of 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—Texanol
ester alcohol. Internal report, Eastman Kodak Company.
Sheldon, L.S., and Naugle, D.F. (1994). Determination of test methods for inte-
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

rior architectural coatings. Revised final report to ICF Inc., USEPA Contact No.
68-D2–0131. Report RTI/5522/042-02 FR. Research Triangle Institute, Research
Triangle Park, North Carolina.
Shields, H.C., and Weschler, C.J. (1992). Volatile organic compounds measured at
a 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 and
latex paint. Environmental Science and Pollution Research, 10(4), 209–216.
Sparks, L.E., Guo, G., Chang, J.C., and Tichenor, B.A. (1999a). Volatile organic
compound emissions from latex paint, part 1: Chamber experiments and source
models. Indoor Air, 9, 10–17.
Sparks, L.E., Guo, Z., Chang, J.C.S., and Tichenor, B.A. (1999b). Volatile organic
compound 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 at
the 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 pollutants
in 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 pollutants
from alkyd and latex paints. Office of Research and Development, Research
Triangle 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.
1080 R. L. Corsi and C.-C. Lin

Verschueren, K. (1996). Handbook of environmental data on organic chemicals. 3rd


ed. 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 emission
profiles for interior latex paints. Proceedings of Indoor Air ‘96, 2, 55–60.
Wolkoff, P. (1998). Impact of air velocity, temperature, humidity, and air on
long-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 adsorptive
reservoir 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 VOC
emission from paints for interior application using test chambers. Proceedings
of Indoor Air ‘99, 5, 155–160.
Downloaded By: [Corsi, Richard L.] At: 04:11 4 December 2009

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

You might also like