You are on page 1of 8

Journal of Polymers and the Environment, Vol. 11, No.

4, October 2003 ( 2003)

A Review of the Fate and Effects of Silicones in the


Environment
D. Graiver,1,3 K. W. Farminer,2 and R. Narayan1

Silicones are well-known useful materials varying in structure, reactivity, and chemical and physical properties, but they all contain a covalent bond between the silicon atom and an organic group.
Most common of these polymers are those based on polydimethylsiloxane (PDMS) having a siloxane (Si-O-Si) repeat unit and two methyl groups on each silicon atom. All these polymers are manmade, and the organosilicon linkage is not found in nature. It was therefore erroneously assumed
that these polymers do not degrade naturally in the environment. It is the purpose of this review to
refute this myth and to describe the degradation processes of PDMS in the environment and any
potential ecological impact on the terrestrial, aquatic, and atmospheric compartments. Although it
was found that minor degradation takes place by hydrolysis of PDMS to dimethylsilandiol followed
by oxidation of the methyl group to aldehyde and ultimately to CO2 by Arthobacter and Fusarium
oxysporium schlechtendahl, the major degradation processes are abiotic. High molecular weight
PDMS are initially depolymerized by soil hydrolysis of the siloxane bonds to yield organosilanol
terminated oligomers. These organosilanols and low molecular weight linear PDMS and cyclics are
evaporated into the atmosphere and are oxidized there by hydroxyl radicals to benign silica, water,
and CO2.
KEY WORDS: Degradation; organosilicon; polydimethylsiloxane (PDMS); silicone hydrolysis; soil; hydroxyl
radicals.

INTRODUCTION

groups from the silicon atom. The purpose of this review


article is to refute this myth and present data that show
that linear silicone polymers and cyclosiloxanes are
readily degraded in the environment by natural
processes. It is further intended here to present data that
show that neither these materials nor their degradation
products pose any harm to the environment or to the living organisms in it. More comprehensive and detailed
articles on this subject can be found in the general literature and in a recent book [1] that also include regulatory status related to these polymers in the United States,
Europe, and Japan at present.
Silicon is the most abundant element in the crust
of the earth after oxygen. However, it is found in nature
exclusively bonded to oxygen in various silicate
structures (e.g., sand, clay, glass, etc.). The most common organosilicon polymer is polydimethylsiloxane
(PDMS). It consists of alternating silicon and oxygen

It is erroneously believed that all silicon-based


polymers, also known as silicones, do not degrade and
persist in the environment for a very long time. Perhaps
this belief is rooted in the fact that organosilicons are
purely manmade materials, and because no biological
process is known that produces a linkage between a silicon atom and a methyl group (or any other organic
group), there are no biological processes that will degrade these linkages and remove an organofunctional

Michigan Sate University, Department of Chemical Engineering and


Material Science, East Lansing, Michigan.
2
BioPlastic Polymers and Composites, Midland, Michigan.
3
To whom correspondence should be addressed. Michigan Sate
University, Department of Chemical Engineering and Material
Science, East Lansing, Michigan. E-mail: dgravier@aol.com

129
1566-2543/03/1000-0129/0 2003 Plenum Publishing Corporation

130

Graiver, Farminer, and Narayan

atoms (e.g., siloxane linkage) as the repeat unit in the


main polymer chain and two methyl groups attached to
each of the silicon atoms. Unlike most synthetic polymers, which are prepared from monomers derived from
petroleum-based oil, PDMS is prepared from a clean
source of silica (e.g., quartz, sand, etc.) and methanol [2].
In this respect, it should be noted that these starting materials are readily available in the environment for the
foreseeable future and, at present, only the energy
needed in the synthetic process is derived from petroleum-based oil.
Although the focus of this review is on cyclic and
linear polyorganosiloxanes, many silanes such as
alkylchlorosilanes, alkoxysilanes, and carboxysilanes are
also included. These organosilanes are commonly used
as intermediates in the polymerization process, as
crosslink agents, process aids, surface treatment agents,
and chemical modiers of other polymers. In general,
such silanes are very susceptible to hydrolysis and thus,
in the presence of moisture, are readily reacted to yield
silanols, which are then condensed to yield low molecular weight organosiloxanes (scheme 1).
In addition to these simple organodisiloxanes, this
review is concerned with the fate and effects of cyclosiloxanes such as octamethylcyclotetrasiloxane, linear oligomers, and high molecular weight polymers, as
well as reactive polyorganosiloxanes. These organosilicon compounds are liquids under ambient conditions
even at high molecular weights. They will flow after
disposal and are therefore more susceptible to migration and movement within the environment than other
solid silicon-based materials such as silicone gums,
elastomers and resins, blends and copolymers of PDMS
with other organic polymers, semiconductor grade silicon and various silicon ceramics, which are outside the
scope of this review. Also not included in this review
are reactive organofunctional silanes that have been especially synthesized to exhibit antimicrobial activity
[3].
Organosilicon polymers are produced by several
corporations and are then sold to a wide spectrum of industries for a variety of applications. For example,
sealants and paint additives are used in the construction
industries; various softeners and foam control additives
are being used in the textile industries; different process

aids and silicone inert ingredients are being used in the


pharmaceutical industry; emulsions containing various
organosilicon compounds are used in personal care, etc.
Consequently, it is not surprising that silicone polymers
enter (at least initially) all three compartments of the environment: atmospheric (air), terrastrial (soil), and
aquatic (water) compartments.

DISCUSSION
Early Work
Early studies concluded that silicon-carbon bonds
are cleaved by biological process because rats that were
fed phenyltrimethylsilane, excreted hydroxyphenyl hydroxymethyl derivatives [4]. However, this study only
showed that silicon-phenyl linkage was cleaved and the
more stable silicon-methyl linkage remained intact. In
another early study cells that were grown on tetraethylorthosilicate (TEOS) and fed dimethyldiethoxysilane as
the carbon source were found to contain traces of inorganic silicate [5]. It was also noted that the oxygen consumption in this case was higher when compared to the
use of ethanol as the carbon source. However, it was argued that the source of the silicate could be from TEOS
and not the organosilane.
In another study a culture of Pseudomonas putida
was grown on PDMS oil, which was used as the carbon
source [6]. Surprisingly, it was concluded that higher
molecular weight PDMS were better growth media,
which implies that higher molecular weight polymers are
more digestible than lower molecular weight polymers.
The most direct evidence of a silicon-carbon cleavage in the environment was published only recently [7],
indicating that 14C labeled PDMS in soil degraded and
yielded 14CO2. The presence of the labeled CO2 was indisputably derived from the labeled C in PDMS.
Subsequent publications have shown that optimum yield
of 14CO2 was obtained when the soil was rst dried to
2% water content followed by rewetting to 14% water;
almost no degradation was observed in soil that was not
allowed to dry [8, 9]. Addition of alfalfa to the soil stimulated biological activity and the production rate of
14
CO2 increased four-fold. In general, the rate of degradation depended upon the water content, pH, tempera-

Scheme 1. General hydrolysis and condensation of organosilanes.

Silicones in the Environment

131

ture, sorption/desorption rate, and availability of oxygen


and other nutrient sources. Based on these results it was
suggested that the rst step in the degradation process is
the hydrolysis of PDMS to dimethylsilandiol followed
by oxidation of the methyl group to aldehyde and ultimately to CO2. The active organisms responsible fo the
oxidation were identied as Arthobacter species and
Fusarium oxysporium (schlechtendahl). These organisms are known to degrade organic compounds by hydroxylation of CH bonds. Similarly, it was argued that
hydroxylation of the methyl group attached to the silicon
atom followed by oxidation would yield aldehyde or carboxylic groups attached to the silicon atom. These functional silanes are hydrolytically unstable and spontaneously will form the corresponding labeled
formaldehyde or 14CO2. However, it was also noticed
that the evaporation rate of the low molecular weight
organosilicons, including the primary dimethylsilandiol
is 220 times faster than the biodegradation rate. This
large difference in rates implies that upon hydrolysis of
PDMS in the soil, most of it is evaporated into the atmosphere, which implies migration of organosilicon
products from one compartment of the environment to
another (terrestrial to atmospheric) and not a complete
degradation. Furthermore, it was noticed that evaporation near the soil surface is predominant over biodegradation, and thus biodegradation is expected to increase
with depth because moisture content is increased.
However, oxygen concentration and rate of hydrolysis
are decreased with depth, which overall reduces the hydrolysis of PDMS to silandiols that are required in the
biodegradation process.
Low Molecular Weight Organosilicons
Based on these early studies, attention has been directed to determine the fate of volatile methylsiloxanes.
These organosilicons include, in addition to dimethylsi-

landiol, other organocyclosiloxanes (most commonly


known as D4, D5, and D6) and relatively low molecular
weight linear oligomers of dimethylsiloxane having a
degree of polymerization up to about ve siloxane repeat
units (molecular weight up to about 400). These low molecular weight organosilicons are used as starting materials for the polymerization of high molecular weight
polyorganosiloxane (gums, uids, and elatomers), as an
inert carrier in personal care applications (e.g., shampoo., moisturizers, antiperspirant carrier, etc.), and also
as a solvent in the electronic industry where they have
replaced chlorofluorosolvents, which were recently
abandoned because of their impact on the ozone layer
The physical properties prole [10] of typical low
molecular weight organosilicons is shown in Table I. It
is evident that these compounds have very limited solubility in water (a few parts per billion), their density is
less than 1.0, and their oil-water separation coefcient
(log Kow) is relatively large. Moreover, they are fairly
volatile for organic compounds of similar molecular
weights because of the very low intermolecular forces
between the siloxane chains and their chemical inertness.
These physical data imply that organosilicons will not
accumulate in the water compartment of the environment. When they are transported into water, they will
quickly phase separate into an oil phase as a lm on top
of the water phase and eventually will evaporate into the
atmosphere. Indeed, data conrmed that these organosilicon compounds pose very low risk to the aquatic
ecosystem as they are found there at very low concentrations [11, 12]; they are transient and rapidly dissipate
from the water through volatilization. Thus these compounds are not a signicant component of the aquatic
compartment, and they do not create any adverse ecological effects in the short time they are present there
[13, 14].
As expected, these low molecular weight, somewhat volatile organosilicons eventually migrate to the at-

Table I. Selected Physical Property Prole of D4, D5, and PDMS

Organosilicon
Water solubility at 25C (ppb)
Vapor pressure at 20C (mm Hg)
Specic gravity (g/cm3)
Aquatic half-life (days)
Log (Kow)
Biodegradation

55
0.7
0.953
16
5.1
Negligible

17
0.12
0.955
12
7.0
Negligible

1.0
Nonvolatile
0.96
NA
10
Negligible

132

Graiver, Farminer, and Narayan

Table II. Reaction Rates of Low Molecular Weight Organosilicons


with Atmospheric Oxidants at 297 2K
organosilicon

O3

NO3

OH

7 1021

8 1017

1.0 1012

7 1021

8 1017

1.4 1012

3 1020

2 1016

5.2 1011

3 1020

2 1016

1.0 1012

3 1020

3 1016

1.5 1012

Note: Reaction rates units are cm3 mole1 sec1.

mospheric compartment of the environment. Their reaction rates with different atmospheric oxidants were
determined [15, 16] and are listed in Table II. It was concluded from these data that the only signicant reaction
of these compounds in the upper atmosphere is oxidation
with hydroxyl radicals. The reactions with NO3 radicals
and ozone are an order of magnitude slower and thus are

insignicant. The lifetime of these compounds from the


values of the reaction rates is very short, and it is predicted that they are completely removed from the atmosphere within 10 to 30 days.
A proposed [17] degradation mechanism is shown
in Fig. 1. According to this mechanism, the methyl group
on the silicon atom is rst attacked by a hydroxy radical
to form a methylene radical (SiCH2) on the silicon atom
and water. Subsequent reaction of this radical with an
oxygen molecule yields a peroxy radical (SiCH2OO).
Because of the high afnity of silicon to oxygen, the peroxy group is rearranged to form an oxycarbon intermediate (SiOCH2O), which in the presence of another
oxygen molecule yields an unstable oxycarbonyl silane
(SiOCH2=O). The oxycarbonlyl silane is readily hydrolized to produce a silanol (SiOH) and various carbonyl compounds, which eventually are oxidized to CO2.
This process is repeated until each of the organic groups
on the silicon atom are cleaved. Condensation of the
silanol groups ultimately leads to silica and water.
Although the exact mechanism may be somewhat different, the important conclusion is that cleavage of the silicon-carbon bond by hydroxy radicals readily occurs in
the upper atmosphere to yield silica, water and CO2.
The impact of low molecular weight organosilicons
on the quality of the air, primarily in urban areas, was
also studied. In particular, it was important to determine
whether these compounds in the lower atmosphere contribute to the formation of tropospheric ozone, smog, or
haze, and whether they have other negative impact of air
quality. Although the atmospheric reactions involving
organic compounds are complicated, many of them involve formation of intermediates that react with NO to

Fig. 1. Proposed atmospheric oxidation mechanism of SiCH3 functional groups.

Silicones in the Environment


form NO2 and organic radicals. A lower concentration of
NO and a higher concentration of NO2 contribute to an
elevated ozone concentration [18, 19]. However, because
volatile organosilicons in the presence of NO2 tend to
terminate free radicals, it was concluded [20, 21] that
even in the presence of other organic compounds, they
have a negative effect on the formation of low-level
ozone. Further studies were conducted in a smog chamber in the presence of known reactive organic gasses to
simulate various urban air quality situations. The results
of these studies clearly indicated that organosilicons do
not contribute to the lower atmosphere aerosol formation
of ozone or smog. Based on these studies the U.S.
Environmental Protection Agency (EPA) excluded
organosilicons from its regulation concerning restriction
of volatile organic compounds in the atmosphere.
High Molecular Weight Organosilicons
High molecular weight organosilicons (molecular
weight between 450 and 450,000) are not volatile and
thus are not expected to transform into the atmospheric
compartment and degrade there like the more volatile,
low molecular weight organosilicons. These high molecular weight polyorganosiloxanes, having primarily a dimethylsiloxane repeat unit, are characterized by a low
melting point far below room temperature and thus,
under ambient conditions, these polymers are owable
liquids even at high molecular weights. It is therefore expected that these polymers will be mobile between the
terrestrial and the aquatic environmental compartments.
High molecular weight PDMS are useful in a wide
range of applications as intermediates in industrial
processes such as release agents, dielectric uids, antifoams, and heat transfer uids. They are also found
uses by individual consumers in various coatings, polishes, detergents, personal care products, food, and pharmaceutics. Consequently, upon disposal, they enter the
environment via different routes. The physical properties
prole (Table I) indicates that these polymers are even
less soluble in water than their low molecular weight
analogue organosilicons and because their density is less
than 1.0 g/cm3, it is expected that even when these polymers enter the aquatic environment, they will phase separate into a lm over the water or rapidly attach to particulates suspended in the water. It is also important to
mention here that because of their high molecular
weight, PDMS do not diffuse through cell membranes
and therefore are not expected to bioaccumulate [22].
Based on wastewater treatment plant models [23, 24] and
laboratory scale calculations [25] it was estimated that
approximately 97% of PDMS eventually ends up in the

133
solid sludge from the waste-water treatment plant.
Actual monitoring of the treated efuent streams from
the wastewater treatment plant also conrmed the absence of PDMS there [26].
Although it is not expected that PDMS would accumulate in the aqueous compartment of the atmosphere
because of its limited solubility and hydrophobic characteristics, PDMS is being transported into the aqueous environment on transit to the wastewater treatment plant.
Various studies on the impact of PDMS during its transient presence in the aquatic environment on microorganisms, invertebrates, amphibians, sh, and other benthic organisms showed no measurable adverse effects.
Furthermore, PDMS had no effect on the wastewater
treatment process parameters (e.g., pH, sludge volume
index, suspended solids and settling data, endogeneous
decay, oxygen uptake rates, half-saturation content for
substrate utilization, and any effect on the aerobic/anaerobic digestion) or the sludge digestion operating parameters [25]. Perhaps the only noticeable effect was reduced foaming during operation, which is not surprising
because PDMS is known for its antifoaming properties
resulting from its very low surface energy. The main
conclusion of these studies was that PDMS does not degrade or decompose in the wastewater treatment process,
it does not affect the wastewater treatment process, and
it ends up almost quantitavely in the solid sludge.
The output solid sludge from the wastewater treatment plant is either incinerated, buried in a landll, or
spread over agricultural elds as fertilizers. Thus, excluding incineration, (which either oxidizes the organic
groups or degrades the polymer to low molecular weight
volatile organosiloxanes that eventually degrade by OH
radicals as described above) the solid sludge from the
wastewater plant ends up in the terrestrial compartment
of the environment.
Several studies on the fate of PDMS in the soil determined that these polymers quickly depolymerize by
soil hydrolysis to break up the siloxane bonds and produce lower molecular weight oligomeric PDMS (Fig. 2)
[7, 27, 28]. Unlike enzymatic cleavage, which is specic
to certain bonds or a depolymerization process from the
chain ends, the soil catalyzed hydrolytic degradation is
random and consequently leads to dramatic reduction in
the molecular weight of the polymer in the very early
stages of the process. Ultimately, this depolymerization
yields predominantly dimethylsilanediol with a minute
amount of trimethylsilanol (derived from the end groups
of the polymer) [29]. These low molecular weight
silanols are fairly volatile and evaporate into the upper
atmosphere, where they are degraded as described above
by OH radicals.

134

Graiver, Farminer, and Narayan

Fig. 2. Proposed soil hydrolytic degradation of PDMS in the soil.

The rate of this depolymerization by hydrolytic


degradation is surprisingly fast. When measuring radioactive PDMS (DP = 195) in aluminum-montmorillonite clays, it was found that after only 30 min the DP
was reduced to 10 and within 21 days all the siloxane
bonds were cleaved with no trace of the polymer (Fig. 3).
Similar depolymerization is observed in other soils; for
example, PDMS incubated in goethite degraded somewhat slower (Fig. 4), but after 30 days no traces of the
polymer in the soil were observed.
In addition to the type of soil, the rate of depolymerization is greatly dependent upon the moisture level.

The fraction of PDMS remaining in the soil (Almontmorillonite) at different water content is plotted in
Fig. 5. It is observed that the rate of degradation is inversely proportional to the relative humidity (RH). thus
depolymerization is much faster at an RH of 32% compared with depolymerization at 100%. The degradation
by depolymerization through cleavage of the siloxane
bonds and the formation of silanols was studied further
to determine the effect of the type of soil, the geographical location, pH, its composition (% organic matter,
sand, silt, and clay), and moisture level [30]. The data indicate that in all these cases PDMS depolymerization occurs within a few days to 3 weeks (Table III). In another
study, forest and agricultural soil were examined and
found to contain no measurable levels of PDMS [31]. In
comparison, soils that were used as meadows, gardens,
and elds (some of them amended with sludge) were
found to contain small quantities of PDMS (0.330.63
mg/kg). However, and this PDMS concentrations were
similar in the treated and untreated soils, and this was
taken as evidence of PDMS degradation in the soil. Soils
that were previously amended with high sludge loadings
(3002880 mt/ha) were also examined and found to have
only very small concentrations of PDMS [32].

Fig. 3. GPC traces of PDMS (initial DP = 195) incubated in


Al-montmorillonite.

Fig. 4. GPC traces of PDMS (initial DP = 195) incubated in goethite.

Fig. 5. Effect of relative humidity on PDMS depolymerization


incubated in Al-montmorillonite.

Silicones in the Environment

135
Table III. Rate of Degradation of PDMS in Various Soils
% Water after drying at

Location

State

PH

% Organic
matter

Appling
Cashmont
Fargo
Hasting
Londo
Tuscola
Wahiawa

GA
WA
ND
NB
MI
OH
HW

5.7
6.8
7.6
5.8
7.6
7.2
4.7

0.9
2.2
5.2
3.0
2.4
0.7
1.9

% Sand

% Slit

% Clay

0.3 bar

air-dried

Degradation
half-life (days)

76
60
2
18
50
69
48

14
26
43
62
28
17
20

10
14
55
20
22
14
32

8.0
13.1
30.7
22.8
17.7
8.9
30.3

1.2
2.3
7.4
3.3
3.1
1.5
4.1

47
710
2128
710
1014
710
47

The conclusion from all these studies is that PDMS


enters the environment by various routes, but most of it
ends up in the soil as part of the sludge from the wastewater treatment plant. In the soil, PDMS undergoes depolymerization through cleavage of the siloxane bonds
to yield low molecular weight silanols. These silanols
are sufficiently volatile to be transported into the atmosphere, where they are degraded by hydroxyl
radicals.
The impact of PDMS and its organosilanol degradation products in the soil was also studied to determine
any adverse effect on micro-organisms or plant growth
[33]. This was done by placing a radioactive-labled
PDMS in the soil and growing successive crops of wheat
and soybeans over a 7-month period in a greenhouse
while monitoring the soil micro-organisms. It was found,
as was mentioned earlier, the PDMS depolymerized to
lower molecular weight oligomers, which were then lost
into the atmosphere. The presence of PDMS and its
degradation products neither affected the plant development nor had any effect on the soil micro-organizms.
The benign effect of PDMS is demonstrated in another
study in which earthworms were placed in highly organic
soil containing 1100 mg PDMS per kilogram of soil [34].
After several weeks it was determined that the presence
of PDMS had no adverse effects on the reproduction of
the earthworms; the total numbers of cocoons, their
mean weight, mean number of earthworms per cocoon
and the mean weight of the earthworms were the same as
in the control soil that contained no PDMS.

CONCLUSIONS
This fate of linear and cyclic polyorganosiloxanes
and their effect on the environment has been studied and
reviewed. It was shown that upon disposal, these polymers enter the environment by many different routes.
However, the low molecular weight oligomers and cyclic
siloxanes are sufciently volatile to be transported to the

upper atmosphere, where the SiC bonds are cleaved by


hydroxyl radicals to yield silica, water, and carbonyl
compounds. High molecular weight polyorganosiloxanes
are transported into the terrestrial compartment of the environment, where they undergo soil hydrolysis in the dry
soil to yield low molecular weight silanol terminated
oligomers. These oligomers are sufciently volatile to
evaporate into the atmosphere and are degraded by hydroxyl radicals.
The rate of degradation is surprisingly fast and implies no accumulation of these polymers or their degradation products in the environment; the degradation rate
in the upper atmosphere is complete in about 1030
days, and hydrolytic degradation in dry soil takes about
47 days. The rate of hydrolysis depends on the type of
soil, pH, organic matter, temperature, and especially the
moisture content.
It was further determined that neither the organosilicons nor their degradation products have any adverse effect on terrestrial and aquatic organisms or plants. The
presence of high molecular weight organosilicons does
not affect the operation of the wastewater treatment plant
and the volatile, low molecular weight oligomers and
cyclics do not affect air quality. The data indicate
organosilicons derived from dimethylsiloxanes do not
pose a risk to the environment.
REFERENCES
1. G. Chandra, ed. (1997) Organosilicon Materials: The Handbook
of Environmental Chemistry, Springer, New York.
2. A. Tomanek (1990) Silicones and Industry, Wacker-Chemie
GmbH, Munich.
3. A. J. Isquith, E. A., Abbot, and P. A. Walters (1973) Appl.
Microbiol. 23, 859.
4. R. J. Fessenden and R. A. Hartman (1970) J. Med. Chem. 13, 52.
5. W. Heinen (1977) Proceeding of the 40th Nobel Symposium,
Sweden, Aug. 2329.
6. R. Wasserbauer and Z. Zadak (1990) Folia Microbiol. 35, 384.
7. R. G. Lehmann, S. V. Varaprah, and C. L. Frye (1994) Environ.
Toxicol. Chem. 13, 1753.
8. R. G. Lehmann, J. R. Miller, and P. Collins (1977) Water, Air,
and Soil Pollution, 106, 111.

136
9. C. L. Sabourin, J. C. Carpenter, T. K. Leib, and J. L. Spivak
(1996) Appl. Environ. Microbiol. 62, 4352.
10. J. Hobson and E. Silberthorn (1995) Environ. Toxicol. Chem. 14,
667.
11. J. Hamelink, P. Simon, and E. Silberhorn (1996) Environ. Sci.
Technol. 30, 1946.
12. D. Kent, P. Fackler, D. Hartley, and J. Hobson (1996) Environ.
Tox. Water Qual. 11, 145.
13. J. Mueller, D. Di Toro, and J. Maeillo (1995) Env. Toxicol. Chem.
14, 1657.
14. D. Kent, P. McNamara, A. Putt, J. Hobson, and E. Silberhorn
(1994) Ecotoxicol. Environ. Safety 29, 372.
15. R. Atkinson (1991) Environ. Sci. Technol. 25, 863.
16. R. Sommerlade, H. Parlar, D. Wrobel, and P. Kochs (1993)
Environ. Sci. Technol. 27, 2435.
17. R. Atkinson, E. Tuazon, E. Kwok, J. Arey, S. Aschmann, and I.
Bridier (1995) Chem. Soc. Faraday Trans. 91, 3033.
18. R. Atkinson (1994) J. Phys. Chem. Ref. Data, Monograph 2, pp.
1216.
19. J. Seinfeld (1996) Atmospheric Chemistry and Physics of Air
Pollution, Wiley, New York.
20. W. P. L. Cartet (1994) Air Waste 44, 881.
21. W. P. L. Carter and R. Atkinson (1989) Environ. Sci. Technol. 23,
864.

Graiver, Farminer, and Narayan


22. R. B. Annelin and C. L. Frye (1989) Sci. Total Environ. 83, 1.
23. E. Namkung and B. E. Rittman (1987) J. Water Pollut. Con. Fed.
59, 670.
24. C. E. Cowan, R. J. Larson, T. C. Feijtel, and R. A. Rapaport
(1993) Water Res. 27, 561.
25. R. J. Watts, S. Kong, C. S. Haling, L. Gearhart, C. L. Frye, and
B. W. Vigon (1995) Water Res. 29, 2405.
26. N. J. Fendinger, D. C. McAvoy, and W. S. Eckhoff (1997)
Environ. Sci. Technol. 31, 1555.
27. R. R. Buch and D. N. Ingebrigstone (1979) Environ. Sci. Technol.
13, 676.
28. J. C. Carpenter, J. A. Cella, and S. B. Dorn (1995) Environ. Sci.
Technol. 29, 864.
29. J. Spivak and B. Dorn (1994) Environ. Sci. Technol. 28, 2345.
30. R. G. Lehmann, S. Varaprah, R. B. Annelin, and J. Arndt (1995)
Environ. Toxicol. Chem. 14, 1299.
31. F. Siebert (1988) Ph.D. Dissertation Ruprecht Karl University,
Heidelberg, Germany.
32. D. C. McAvoy, K. M. Kerr, and N. J. Fendinger (1996) Abstract
No. 254, SETAC 17th Annual Meeting, Washington, DC.
33. R. G. Lehmann, C. L. Frye, D. A. Tolle, and T. C. Zwick (1996)
Water Air Soil Pollut. 87, 231.
34. N. Garvey, M. K. Collins, and E. M. Mihaich (1996) Abstract No.
256, SETAC 17th Annual Meeting, Washington, DC.

You might also like