You are on page 1of 190

Springer-Verlag 2004

Rev Environ Contam Toxicol 182:1195

Photodegradation of Pesticides
on Plant and Soil Surfaces
Toshiyuki Katagi
Contents
I. Introduction .......................................................................................................
II. Photophysical and Photochemical Processes ...................................................
A. Photophysical Processes ..............................................................................
B. Photochemical Processes .............................................................................
III. Factors Controlling Photolysis on Plant Surfaces ...........................................
A. Environmental Factors .................................................................................
B. Illumination Conditions ...............................................................................
C. Effect of Formulation ..................................................................................
D. Anatomy of the Leaf ...................................................................................
E. Wax Chemistry .............................................................................................
F. Photoinduced Reactions ...............................................................................
IV. Factors Controlling Photolysis on Soil Surfaces .............................................
A. Soil Components ..........................................................................................
B. Environmental Factors Affecting Soil Properties .......................................
C. Mass Transport in Soil .................................................................................
D. Photic Depth in Soil ....................................................................................
E. Effects of Soil Properties on Photolysis ......................................................
F. Photophysical and Photochemical Processes of Soil Components .............
V. Atmospheric Oxygen Species ..........................................................................
VI. Experimental Design and Kinetic Analysis .....................................................
A. Light Source .................................................................................................
B. Photolysis Chambers ....................................................................................
C. Kinetic Analysis ...........................................................................................
VII. Photodegradation of Pesticides in Model Systems ..........................................
A. Soil Surface Models ....................................................................................
B. Plant Surface Models ...................................................................................
C. Photodegradation of Pesticides on Glass and Silica Gel Surfaces .............
D. Photodegradation of Pesticides in Organic Solvents and
Plant Model Systems ...............................................................................
VIII. Photodegradation of Pesticides on Soil and Clay Surfaces ............................
IX. Photodegradation of Pesticides on Plants ........................................................
Summary ....................................................................................................................

2
3
3
5
10
10
10
12
13
14
15
17
17
19
19
21
22
23
29
31
31
32
34
35
35
36
36
47
56
69
77

Communicated by George W. Ware


T. Katagi
Sumitomo Chemical Co., Ltd., Environmental Health Science Laboratory, 2-1 Takatsukasa
4-Chome, Takarazuka, Hyogo 665-8555, Japan.

T. Katagi

Table Listing .............................................................................................................. 79


Appendices ................................................................................................................. 129
Directory of Pesticide Chemical Structures .............................................................. 130
References .................................................................................................................. 157

I. Introduction
Sunlight photodegradation is one of the most destructive pathways for pesticides
after their release into the environment. Plant surfaces, especially leaf surfaces,
are the first reaction environment for a pesticide molecule after application, and
spray drift would indirectly present a similar situation. Photolysis on soil surfaces becomes important when a pesticide is directly applied to soil or not significantly intercepted by plants, providing that the leaf cover does not shade the
ground from sunlight. Because the foliar interception of pesticides depends on
plant species and usually increases with their growth stage (Linders et al. 2000),
the importance of soil photolysis is considered to be lessened when plants become mature. Spray drift after pesticide application or washoff from plants by
rain is the indirect route by which a pesticide reaches the soil.
To elucidate the photodegradation profiles of pesticide in the environment,
many investigators have focused on solution photolysis in organic solvents or
in a dilute aqueous solution. The heterogeneity of soil and plant surfaces together with the capricious transmission of sunlight onto these media also makes
photolysis on them more difficult to understand. Although there are many excellent reviews on photolysis of pesticides (Roof 1982; Miller and Zepp 1983;
Choudhry and Webster 1985; Marcheterre et al. 1988; Parlar 1990; Wolfe et al.
1990; Cessna and Muir 1991; Meallier 1999; Floesser-Mueller and Schwack
2001; Burrows et al. 2002), photolysis on soil is still only partially understood
because of the limited number of investigations, whereas that on plants is mostly
speculation derived from plant metabolism studies. Under these circumstances,
photodegradation on soil and plant surfaces requires more examination, not only
experimentally but also theoretically, to reveal the mechanisms controlling photophysical and photochemical processes in pesticides on solid phases and to
apply such knowledge to better understand the dissipation profiles of pesticides
in the field.
This review first considers the background and basis of photophysics and
photochemistry relevant to the photodegradation of pesticides. Molecular excitation and deactivation processes together with subsequent chemical processes are
discussed, and reactivity toward active oxygen species is briefly summarized.
Second, constituents of soil and plant surfaces are reviewed from the point of
view of the factors controlling photodegradation, together with meteorological
factors. After reviewing the experimental design of photodegradation on these
surfaces, the analysis of experimental data in consideration of the photodegradation mechanism is discussed briefly. Based on the literature survey, both model
systems and the actual photodegradation in soil and plant systems are reviewed

Photodegradation of Pesticides

for every chemical class of pesticide. The chemical structure of each pesticide1
appearing in this review, together with a corresponding number in bold type, is
provided in the Appendices.

II. Photophysical and Photochemical Processes


A. Photophysical Processes
The extent of sunlight photolysis is highly dependent on UV absorption profiles
of the pesticide, the surrounding medium, and the emission spectrum of sunlight.
Because the energy to break chemical bonds in pesticide molecules usually
ranges from 70 to 120 kcal mol1, corresponding to light at wavelengths of
250400 nm (Watkins 1974), spectral irradiance of sunlight detected near the
ground becomes important in determining the photodegradation profiles of pesticide. By passing through the atmosphere, sunlight intensity significantly decreases to about 10% in the troposphere, and no light is transmitted at wavelengths from <290 to 295 nm, mainly due to absorption by ozone (Zepp and
Cline 1977; Parlar 1990). As a result, sunlight near the ground exhibits a maximum at around 440460 nm, and its intensity at the UV region responsible
for photodegradation of pesticide becomes approximately 5%6% of the total
intensity. There are many photophysical pathways of sunlight absorption (Fig.
1) (Turro 1978; Roof 1982; Parlar 1990). When a photon passes close to a
pesticide molecule, molecular excitation occurs via interaction between the electric field of a pesticide molecule and that of light at a time scale of femtoseconds
without a change of molecular geometry (FranckCondon principle). Each photon can activate only one molecule in the ground state (S0) with a certain probability to the excited singlet state (StarkEinstein rule), and usually the lowest
excited state (S1) is involved in further photoprocesses. Generally, pesticide molecules exhibiting a UV-vis absorption spectrum at >290 nm have a substituted
aromatic moiety, sometimes being conjugated with the lone-pair electrons or the
unsaturated bonds such as carbonyl or carbamoyl group, and hence * or
n * transition takes place upon irradiation.
There are three possible photophysical pathways from the S1 state: nonradiative internal conversion, emission of fluorescence, and intersystem crossing to
the excited triplet state (T1) (see Fig. 1). The first pathway means the relaxation
from higher vibrational levels (1012 sec1) in the S1 state followed by decay to
a lower electronic state with the same multiplicity (106 1012 sec1). The second
one is a radiative deactivation process. The fluorescence spectrum is usually
close to a mirror image of that of absorption due to the FranckCondon principle but shifted to the red. The lifetime of fluorescence is very short (nanosec-

All pesticides are identified by their common name and parenthetical identification number, and
their structures are illustrated in the Appendices. Preceding the Appendices is a Directory of Pesticide Chemical Structures listing all pesticides in alphabetical order to aid the reader in locating
specific structures.

T. Katagi

Fig. 1. Energy state diagram.

onds to microseconds) due to the transition between states with the same multiplicity. The last pathway is a spin-forbidden process (S1 T1), followed by
slow radiationless deactivation or emission of phosphorescence. The T1 S0
process is also spin-forbidden, and hence the lifetime of phosphorescence usually becomes an order of milliseconds to 102 sec. The profiles of fluorescence
and phosphorescence spectra of pesticides, based on the literature survey, are
summarized in Table 1 (see table on page 80). Although there are many chemical classes and either a solvent system or temperature difference in measuring
spectra, their maximum wavelengths are located in the range of 280450 and
380530 nm, respectively. Based on the following equation of conversion where
E and are the energy level and the emission wavelength, respectively (Gould
1989b), the energy levels of the excited singlet (Es) and triplet (ET) states can be
estimated to be 64102 and 5475 kcal mol1 for these pesticides, respectively.
E (kcal mol 1) = 2.864 104/ (nm)
Because intersystem crossing is facilitated by the presence of heavy atoms in a
molecule, the fluorescence spectrum of a pesticide is usually difficult to measure
at room temperature, but in such cases phosphorescence can be efficiently detected instead.
The foregoing consideration can also be applied to pesticide molecules in the
solid phase, but adsorption onto these media is most likely to affect the photophysical processes. Molecular motion would be highly restricted, and interactions with these heterogeneous surfaces result in modification of their electronic
states. In this case, the reflectance spectrum of a pesticide gives more useful
information than an absorption spectrum, and this is described by the relation-

Photodegradation of Pesticides

ship of Schuster and Kubelka-Munk (Parlar 1984) instead of the BeerLambert


law.
F(R) = (1 R)2/2R = K/S
The diffuse reflectance, F(R), represents the radiation penetrating into the powder and resembles the usual transmission spectrum. R is the ratio of reflectance
of a sample to that of a standard and thus the relative diffuse reflectance of an
infinitely thick layer compared to a nonabsorbing standard such as magnesium
oxide; K and S are the absorption and scattering coefficients, respectively. Adsorption can produce unequal displacement of the ground- and excited-state potential curves, which would result in a different vibronic band shape. Thus,
spectral changes by adsorption are characterized by a spectral shift, changes of
extinction coefficient, broadening of absorption bands, and appearance of new
bands (Wendlandt and Hecht 1966; Nicholls and Leermakers 1971; Parlar
1984). Examples of spectral changes by adsorption of organic molecules including pesticides to silica gel and clays, listed in Table 2 (see table on page 85),
are based on the literature survey. Both bathochromic (red) and hypsochromic
(blue) shifts on adsorption have been reported, which are considered to depend
on the type of an electronic transition. It is known that a blue shift almost always
occurs with n * transition and often a red shift with * transitions
(Nicholls and Leermakers 1971). In the case of the former transitions, the
change of a nonpolar environment to polar causes more stabilization of the
ground state via hydrogen-bonding and dipoledipole interactions than the excited state, resulting in a blue shift. For * transitions, the excited state is
more stabilized by polarization in the polar environment, resulting in a red shift.
The alteration of emission spectrum by adsorption is likely, but the corresponding information is limited. Villemure et al. (1986) have reported the significant
increase of fluorescence intensity of paraquat (225) when adsorbed onto clays.
Fluorescence with an emission maximum of 345 nm was very weak in aqueous
solution, but adsorption resulted in the increase of its intensity with a blue shift
by 20 nm. The increase of intensity is most likely to stem from an inhibition
of radiationless quenching by counteranion Cl by intercalation of molecules of
(225) into the interlayer of clays.
B. Photochemical Processes
Unless the energy of an excited-state molecule is lost as heat or emission of
light, it causes various types of chemical reactions in the excited molecule.
There are two types of photochemical reactions, well known as direct and
indirect photolysis (Roof 1982; Miller and Zepp 1983). Direct photolysis
means the photoreaction proceeds by absorbing light energy, whereas indirect
photolysis is defined as reaction of a ground-state molecule with the other excited molecule or photochemically produced reactive species. The former indirect photolysis is called photosensitization or quenching, and the latter is a photoinduced reaction with a reactive oxygen species. The average rate of direct

T. Katagi

photolysis in a well-mixed system can be estimated by using the GCSOLAR


program based on spectral irradiance of sunlight, absorption profiles, and quantum yield of pesticide (Leifer 1988). In contrast, when pesticide molecules exist
as deposits on soil and plant surfaces, the heterogeneous microenvironment
makes such estimation difficult. For example, many researchers have reported
the quantum yield for pesticides in solution photolysis, but the information is
very limited on solid-phase photolysis (Krieger et al. 2000; Samsonov and Pokrovskii 2001). In the case of soil photolysis, Balmer et al. (2000) introduced a
model function of light attenuation in soil with diffusion of a pesticide molecule
to better describe the dissipation profiles.
The molecule in the S1 or T1 state undergoes various chemical reactions.
Typical reactions observed in photolysis on soil and plant surfaces are summarized in Fig. 2. One of the most important photoreactions is initiated by carbonyl
n *excitation. The photoinduced cleavage of a CC bond generates the
ketyl radical (Norrish type I), or the carbonyl carbon in the excited state abstracts hydrogen from a neighboring alkyl group (Norrish type II) or a solvent
molecule. The electronic excitation also occurs at the C=C bond or aromatic
moiety, which results in cis/trans (or E/Z) geometric isomerization or R/S optical
one. The photoinduced homolytic bond cleavage is also a main reaction pathway
in photolysis. When it occurs at an ester or ketone moiety, decarboxylation or
decarbonylation proceeds in addition to the apparent photoinduced hydrolysis,
and its extent depends on solvent structure in relation to stabilization of produced radicals. The C-halogen (Cl, Br, and I) bond is known to also undergo
photoinduced cleavage. Dealkylation via oxidation with O2 or reactive oxygen
species such as the hydroxyl radical (OH) is also known. Oxidation at either
carbon or sulfur is one of the most important routes of photodegradation. The
most familiar rearrangement is thiono-thiolo for O-aryl phosphorothioate insecticides whose O-alkyl group (typically a methyl group) shifts to the P=S sulfur
atom (reaction 9a). The other is a photo-Fries rearrangement for amides and
carbamates where the ketyl radical generated via cleavage of an NC(=O) bond
mostly migrates to the o- or p-position of the phenyl ring (reaction 9b). The
formation of a new bond is typically observed for intramolecular cyclization for
organochlorine cyclodiene insecticides (reaction 10a).
Incidentally, an energy transfer can proceed between the excited donor (D*)
and acceptor (A) molecules. The spectral overlap between the emission spectrum of D* and absorption spectrum of A is prerequisite (Fig. 3), and energy
transfer proceeds efficiently when the process is spin-allowed and exothermic
(Turro 1978; Roof 1982). There are two mechanisms known for energy transfer,
coulombic and exchange interactions. The former mechanism involves the induction of a dipole oscillation in A by D* via a magnetic field and does not
require physical contact of the interacting D* and A. Forester theory indicates
that the rate of energy transfer according to this mechanism is proportional to
the spectral overlap and inversely proportional to intermolecular distance between D* and A to the sixth power. Therefore, the energy transfer efficiency is
and thus is
greatly reduced as the distance increases up to approximately 50 A

Photodegradation of Pesticides

Fig. 2. Types of photoreactions.

sensitive to diffusion of D* and A. In contrast, the exchange interaction involves


a double electron substitution, that is, jump of an electron from D* to the unoccupied orbital of A and the simultaneous jump of an electron from A to the
half-occupied orbital of D* via overlap of electron clouds, which requires physical contact (collision) between D* and A. Singlet energy transfer is spin-allowed
for both long-range coulombic and short-range exchange mechanisms. However,

T. Katagi

Fig. 2. Continued.

because an acceptor molecule is in the S0 state, the triplet energy transfer is


basically spin-forbidden and only proceeds via a short-range exchange mechanism. The interatomic distance expected for the triplet energy transfer is esti . The longer a molecule remains in an excited state the
mated to be 1015 A
greater the probability that it will transfer energy to a suitable neighboring molecule. Therefore, the triplet energy transfer is the most common and most important type of energy transfer involved in photolysis of pesticides.

Fig. 3. Schematic description of the spectral overlap.

Photodegradation of Pesticides

When photolysis on soil and plant surfaces is considered, diffusion of pesticide molecules is likely to be limited by adsorption and very high viscosity of
the medium. These situations may imply less possibility of energy transfer except for the case where pesticide molecules are located in the neighborhood of
D* or A. There are many candidates in the environment playing a role as D* or
A, and some of them are listed in Table 3 (see table on page 86) together with
synthetic chemicals. Because flavonoids and long-chain alkyl ketones are some
of the wax components in plant foliage (see Section III.E), photosensitization
by these components may be possible by taking account of the ET values of
pesticides (5475 kcal mol1). In the case of soil surface, either photosensitization or quenching by humic substances with the ET value of 6062 kcal mol1
is considered to proceed. Although concrete demonstrations by measurement are
limited, the importance of spectral overlap between pesticide and synthetic dyes
coadsorbed on clay surfaces has been reported in relation to stabilization of
photolabile pesticides (Margulies et al. 1988; El-Nahhal et al. 2001).
Among natural products, molecular oxygen (O2), whose ground state is a
triplet, is the most effective quencher. The very low lying singlet states with
approximate energy levels of 23 and 38 kcal mol1 can easily react with the
excited states of pesticides and natural products, resulting in the formation of
singlet oxygen (1O2). In addition, the other active species such as OH and ozone
(O3) are deeply involved in photoinduced reactions. It is not easy to identify
active oxygen species in photolysis on soil and plant surfaces, but the basic
reactivity of some pesticides is known (Table 4 [see table on page 88]). Sulfur
oxidation is one of the characteristic reactions. Concrete evidence on the
involvement of 1O2 was given for fenitrothion (138)2 by Verma et al. (1991),
who showed the significant decrease of oxon formation when the 1O2 scavenger
was added. Formation of peroxide at the isobutenyl moiety of pyrethroid (Fig.
2, reaction 7) was found sensitive to 1O2 scavenger (Ruzo et al. 1980, 1982) and
was greatly reduced by introduction of halogen atoms instead of the geminal
methyl groups (Holmstead et al. 1978a; Ruzo 1983). Hirahara et al. (2003)
confirmed the photoinduced formation of 1O2 in the phosphate buffer solution
of fenthion (143) without any dye by ESR (electron spin resonance) using the
spin trap reagent and supposed that (143) is a photosensitizer for O2. Several
methods, including Fentons reagent and illumination in the presence of hydrogen peroxide (H2O2), O3, Fe3+, humic substances, or a semiconductor, have been
utilized to generate OH. The oxidative desulfuration and N-dealkylation together with hydroxylation proceeded via reactions with OH. The involvement
of OH in photolysis of atrazine (185) was demonstrated by retardation of the
reaction in the presence of mannitol as a radical scavenger, and the attack at the
-position of the N-ethyl moiety was evidenced by formation of the N-acetyl
derivative. Concerning O3, aqueous ozonization has been extensively investigated (Reynolds et al. 1989) but the reaction on solid surfaces seems to be

See footnote 1, p. 3.

10

T. Katagi

limited. Spencer et al. (1980) reported the desulfuration of parathion (135) on


soil dusts and clay minerals in the presence of O3, which was recently confirmed
by Kromer et al. (1999). The sulfur atom was finally oxidized to sulfate ion
(Gunther et al. 1970). In the case of pyrethroids, ozonization of the isobutenyl
moiety was found to proceed to give the corresponding aldehyde derivatives
(Ruzo et al. 1982).

III. Factors Controlling Photolysis on Plant Surfaces


A. Environmental Factors
A number of factors such as meteorological conditions, formulation type,
sprayer characteristics, and affinity of plant surface to formulation are considered to determine the amount of pesticide attached to the surface as well as
ground cover and canopy thickness of plants (Willis and McDowell 1987).
Zongmao and Haibin (1997) extensively investigated factors controlling dissipation from tea plant surfaces for 16 pesticides. Photodegradation was found to be
one of the most important factors in dissipation process except for evaporation,
rainfall elution, and growth dilution. Both photolysis and rainfall elution were
found to play a great role in the dissipation of diflubenzuron (159) in a conifer
forest (Rodriguez et al. 2001). Garau et al. (2002) examined the extent of pesticide loss from a cellulose membrane due to evaporation and codistillation in
the presence or absence of underlying water. Evaporation, codistillation, and
photolysis all contributed to dissipation of pyrimethanil (209) and cyprodinil
(210) but with each varying to some extent, while only photolysis was the controlling factor for azoxystrobin (244) and fludioxinil (208). The existence of
tomato fruit wax mostly retarded evaporation and codistillation of pesticides and
exhibited a screening effect against sunlight. For a pesticide with higher vapor
pressure and less photoreactivity, volatilization loss became predominant in dissipation as observed for chloropyrifos (145) (Meikle et al. 1983). Through a
glass wind tunnel study for 14C-fenpropimorph (227) individually applied to
bean, sugar beet, and radish, the importance of reactive species (OH and/or O3)
in air was demonstrated (Ophoff et al. 1999). Furthermore, either soil dusts or
clay minerals enhanced oxidation of parathion (135) to its oxon in the presence
of O3 (Spencer et al. 1980). In addition to these factors, penetration of pesticide
into cuticle and biotic metabolism therein are also considered important (Bentson 1990; Katagi and Mikami 2000).
B. Illumination Conditions
Spectral irradiance of sunlight at the plant surface is most important to understand the effect of photolysis (Fig. 4). Because the window glass used in ordinary greenhouses absorbs a considerable amount of light in the UV-B region

All pesticides are identified by their common name and parenthetical identification number, and
their structures are illustrated in the Appendices. Preceding the Appendices is a Directory of Pesticide Chemical Structures listing all pesticides in alphabetical order to aid the reader in locating
specific structures.

Photodegradation of Pesticides

11

Fig. 4. Photodegradation of pesticide on plant: (a) precipitation, (b) wind, (c) volatilization, (d) sunlight outdoors, (e) sunlight in the borosilicate glass greenhouse.

(280320 nm), this filtering effect is likely to reduce the overlap between the
solar emission spectrum and the near-UV absorption spectrum of many pesticides (Kleier 1994). Photodegradation was measurably reduced by covering the
Petri dish as a model of the greenhouse window (Garau et al. 2002). Fukushima
et al. (2003) examined the photolysis of 14C-fenitrothion (138) on tomato fruit
in a greenhouse with a ceiling made of quartz or borosilicate glass. The intensity
of sunlight at <360 nm was significantly reduced in the borosilicate glass greenhouse, and neither the corresponding oxon nor the S-isomer generated by photolysis in the quartz greenhouse was detected. Furthermore, transmission through
the greenhouse window is also known to be reduced by glass pollution, and its
extent was larger in the shorter wavelength region (Van Koot and Dijkhuizen
1968). The structure of greenhouse changing the intensity and spectral irradiance
of the transmitted sunlight gave an insignificant effect on dissipation of chlorpyrifos (145). Type of crop and season were the most relevant factors (Martnez
Vidal et al. 1998). Similar results were obtained for fenpropathrin (24) (Martnez
Galera et al. 1997), while degradation of methomyl (70) was found to depend
on the type of greenhouse (Gil Garcia et al. 1997).
Degradation of pesticides in the greenhouse or outdoors was compared to
examine the controlling factors in foliar dissipation. The application of pirimi-

12

T. Katagi

carb (78) to lettuce was conducted both in greenhouse and field (151; Cabras et
al. 1990). No significant differences occurred in half-lives of total carbamates,
but greater formation of these degradates was observed in the field. The comparative degradation study of parathion (135) using a growth chamber, greenhouse,
and open field with and without motorized covering exhibited more formation
of the oxon and S-isomer in the field (Joiner and Baetcke 1973). Based on these
results, the experimental conditions of growing plants should be monitored and
compared with the real environment as much as possible to investigate the most
realistic pesticide photodegradation process.
C. Effect of Formulation
Pesticide formulation is composed of an active ingredient, carrier such as clay,
surfactants as wetting and spreading agents, nonevaporating viscous stickers,
humectants, and penetrating agents such as crop oils (Hazen 2000). These additives having hydrophobic and hydrophilic parts in a molecule provided a very
complex medium for photolysis of pesticides, and their aromatic moiety becomes
a possible photosensitizer or quencher (Nutahara and Murai 1984; Thomas and
Harrison 1990). Baker et al. (1983) investigated extensively the changing nature
of epicuticular waxes on the impact of several formulations containing 14Clabeled pesticide using scanning electron microscopy, X-ray analysis, and microautoradiography. Oil formulations were found to immediately spread through
crystalline wax whereas aqueous solutions distributed most readily over smooth
surfaces. Lipophilic pesticides are partitioned favorably into the organic phase,
separating as a zone at the outer edge of the droplet residue, but hydrophilic
pesticides are concentrated in the central region. Furthermore, the fluidity of
epicuticular waxes is known to change with hydrophobicity and lipophilicity
balance of the surfactant (Hess and Foy 2000). In addition, surfactants in formulations are considered to affect either uptake of pesticide molecules across the
cuticle to plant tissues or photodegradation profiles on plant surfaces. The former phenomenon by monodispersing alcohol ethoxylates has been demonstrated
for several pesticides on barley leaves. Larger effects on diffusion coefficient
were observed for the larger-size molecule (Burghardt et al. 1998). The latter
effect was first pointed out in aqueous solution by Tanaka et al. (1979, 1981).
Instead of formation of 4-OH and N-CHO derivatives, monuron (52) in aqueous
solutions of surfactants Tergitol TMN-6 or Triton X-100 underwent dechlorination followed by dimerization and N-demethylation. Furthermore, photodegradation of 17 herbicides was found to be accelerated by the presence of these
surfactants.
Based on these results, pesticide molecules were thought to be partitioned to
hydrophobic cores of micelles where photolysis such as that in an organic solvent proceeded favorably and surfactants such as Triton X-100 having an aromatic moiety could act as a photosensitizer. Such photosensitization was also
reported when oxysorbic or plant oil concentrate was used as the surfactant
(Harrison and Wax 1985). In contrast, it is supposed that some surfactants hav-

Photodegradation of Pesticides

13

ing a lower triplet excited energy than that of the pesticide can act as a quencher,
which has been demonstrated for TMN-6 and nonaethoxylated p-(1,1,3,3-tetramethylbutyl)phenol (Tanaka et al. 1986, 1991). The other possible effect by
surfactants would be stabilization of photoproduced radicals in a cage. In the
case of 2,4-D (1), a few products formed via Norrish type II or photo-Fries
mechanism were detected through photolysis in aqueous solution containing surfactant (Que Hee et al. 1979). There was no significant effect of Tween 80 or
Trion X-100 on photolysis of metsulfuron (97) and chlorimuron (101) on glass
surfaces (Thomas and Harrison 1990), whereas in the other study accelerated
photolysis was observed (Harrison and Thomas 1990). On corn leaves, similar
acceleration was observed for (97) but with no clear effect for (101). In pyrethroids, reduced photolysis on glass surfaces was reported when surfactants
where included (Megahed et al. 1987).
D. Anatomy of the Leaf
As shown in Fig. 5, leaves are covered with protective cuticles that function by
decreasing water loss and protecting the plant from infection by various pathogens. The cuticle is a complex structure consisting of a pectin layer that binds
the cutin to the epidermal cell walls and a layer of epicuticular wax on the
outside, this structure is known to depend on plant species (McFarlane 1995;
Bianchi 1995). When the stomata are open, gas molecules can diffuse in and
out and interact with a large hydrophilic area of water-covered mesophyll cells.
Most pesticides are hydrophobic molecules, and thus the large lipid-covered
surface of leaves (cuticles) forms an ideal sink for accumulation of pesticides.
The fine structure of the wax layer greatly differs between plant species and is
morphologically classified by using light microscopy into four main forms: needles, rods, granular layers and films (Baker 1982). Use of the electron microscope has revealed that the aerial surfaces of all higher plants carry a partial or
continuous coverage of amorphous wax and that formation of crystalline wax is

Fig. 5. Transverse view of the typical surface structure of plant foliage: (a,d) epidermal
cell, (b) stoma, (c) mesophyll, (e) pectin, (f) cutin and embedded waxes, (g) epicuticular
waxes.

14

T. Katagi

frequently superimposed on amorphous layers. Penetration through these wax


regions and the underlying cutin layer has been extensively studied, for example, by using the diffusion cell method (Schonherr and Riederer 1989).
Radiant energy of sunlight is considered to interact with the leaf structure by
absorption and scattering. As shown in Fig. 5, most leaves have a distinct layer
of long palisade parenchyma cells in the upper part of the mesophyll and more
irregularly shaped spongy parenchyma cells in the lower part. Sunlight is reflected and scattered by hairs, leaf pubescence, and the glaucous leaf surface,
and a portion of the light enters into the leaf (Robberecht and Caldwell 1980;
Holmes and Keiller 2002). This light is critically reflected internally at the cell
walls in the intercellular space as a result of the difference of refractive index
between air and water in tissues (Gates et al. 1965). Pesticides by foliar application are considered to distribute mainly on the epicuticular wax layer, but a
portion may enter into the plant directly through stomata or diffusion; thus,
depth and spectral distribution of penetrated sunlight would be important when
photodegradation is considered. Many studies have been conducted to investigate this using a fiberoptic probe (Vogelmann and Bjorn 1984). About 90% of
the penetrating monochromatic light (310 nm) was attenuated within the initial
one third of the leaf (100150 m) of Brassica napus L., mostly at the epidermal cells; polychromatic radiation (280320 nm) exhibited a relatively uniform
spectral distribution within the leaf (Bornman and Vogelmann 1991; Cen and
Bornman 1993). UV-B radiation was found to reach the epidermis and meso lenius et al. 1995). Day et al. (1994)
phyll in other measurements for this leaf (A
measured UV absorption spectra and the epidermal transmittance spectra at
280350 nm of foliage from 42 plant species and demonstrated that some flavonoids act as a UV-absorbing agent. These observations imply that pesticide molecules in the leaf can absorb some part of the radiation energy of sunlight irrespective of their location, and not only the anatomy of the leaf but also
chromophores in leaf tissue can greatly affect their photodegradation.
E. Wax Chemistry
Epicuticular waxes are basically aliphatic compounds and are readily solubilized
by organic solvents with minimal contamination by lipids from the inner cuticular layers. A mixture of chloroform and diethyl ether (1 : 1, v/v) was found to
be efficient to isolate waxes containing cyclic compounds (Baker 1982). Leaves
of many herbaceous plants carry delicate membranes with only sparse wax deposits (510 g cm2) and many weed species also have thin wax layers. Wax
deposits on rapidly expanding leaves of leek and Brassica spp. are heavier
(3060 g cm2), similar to those on leaves of many fruit crops. Wax layers on
fruits are invariably much thicker than those on the corresponding leaves (50
400 g cm2), and thick deposits were also found for pistachio and olive (60
300 g cm2). As a result, the average thickness of the cuticle varies with plant
species from 3 to 15 m (Lendzian and Kerstiens 1991). Wax chemistry, especially of epicuticular waxes, has been systematically investigated for many plant

Photodegradation of Pesticides

15

species, sometimes with different growth stages, first by thin-layer chromatography (TLC) and gas chromatography (GC) of the derivatized extracts and later by
gas chromatography-mass spectrometry (GC-MS), nuclear magnetic resonance
(NMR), and infrared (IR) (Baker 1982; Walton 1990; Bianchi 1995). Waxes are
basically classified into even- and odd-carbon-numbered straight chain homologues and cyclic compounds (Baker 1982; Bianchi 1995). The first class consists dominantly of acids (C12 C18 and C20 C32), aldehydes (C22 C32), primary
alcohols (C22 C32), and alkyl esters (C36 C72). The odd-numbered homologues
are hydrocarbons (C17 C35), secondary alcohols (C21 C33), ketones (C23 C33),
-diketones (C29 C33), and hydroxy--diketones (C29 C33). The last class consists of triterpenoid acids such as ursolic and oleanolic acids, triterpenols such
as - and -amyrin, triterpenoid esters, and ketones. Among them, the following
chemicals are known to be common major constituents of epicuticular waxes:
nonacosane and hentriacontane for hydrocarbons: hexacosanol, octacosanol, and
triacontanol for primary alcohols; nonacosan-10-ol for secondary alcohols; hentriacontane-14,16-dione and tritriacontane-16,18-dione for -diketones; and ursolic acid for triterpenoids.
Typical wax constituents with major homologues for representative plants,
fruits, and leaves are briefly summarized in Table 5 (see table on page 89). In
the case of leaf waxes, composition was found to vary with not only growth
stage but also the parts of the leaf, that is adaxial and abaxial surfaces (Bukovac
et al. 1979; Baker and Hunt 1981; Riederer and Schneider 1990). Unique components such as anthraquinone in the leaf waxes of perennial rye grass (Allebone
et al. 1971), a long-chain 1,4-benzoquinone in a wide variety of leaves (Kofler
et al. 1959), and novel fatty acid esters of E- and Z-p-coumaryl alcohols in cv.
Gala apples (Whitaker et al. 2001) have been reported.
F. Photoinduced Reactions
There has been no systematic investigation on the contribution of these wax
components to photolysis of pesticides. Pirisi et al. (1998) have measured UV
absorption spectra of epicuticular waxes of nectarines, oranges, and mandarin
oranges extracted by chloroform and/or methanol. These waxes basically exhibited featureless absorption at 290400 nm with small shoulders around 320 nm,
but no clear correlation with photolysis rates of pirimicarb (78) could be detected. The wax of annual spruce needles also showed the featureless spectrum
but with significant absorption at 300400 nm (Niu et al. 2003). Furthermore,
although no absorption maximum was observed for leaf waxes of caster bean,
UV absorption at about 300 nm was observed for Scotch pine, cabbage, and
Carnauba waxes (Wuhrmann-Meyer and Wuhrmann-Meyer 1941). Concerning
each chemical class of waxes, the theory of UV absorption indicates no absorption by simple hydrocarbons, but many chemicals containing a carbonyl moiety
can absorb UV light at >290 nm (Jaffe and Orchin 1962). n-Hentriacontane, the
major component of green tobacco leaf, was shown to be transparent to the UV
region (Carruthers and Johnstone 1959). As -diketones and ,-unsaturated

16

T. Katagi

ketones (aldehydes) show UV absorption maxima around 270 nm and 320370


nm (Cookson and Dandegaonker 1955; Horn and Lamberton 1962), these components as well as anthraquinone and 1,4-benzoquinone already described are
most likely to be involved in photolysis of pesticides on plant foliage.
Triterpenoid derivatives are detected as major to dominant components of
leaf waxes, but their UV absorption maxima are usually located at 200210 nm
in hexane (Wheeler and Mateos 1956; Weizmann and Mazur 1958; Turner
1959), which may indicate less importance in photolysis of pesticides. Phenylpropanoids including flavones, flavonols, and cinnamoyl esters in higher plants
are known to exhibit UV screening effect (Cockell and Knowland 1999; Kolb
et al. 2001). These flavonoids are also known to show an antioxidant activity
(McPhail et al. 2003) similar to -diketones (Osawa 1994). Because the complex
mixture of wax components is the primary reaction medium for pesticides, it
should be kept in mind whether these components simply act as highly viscous
organic solvents or are involved in photochemistry of pesticides as photosensitizer or quencher. As an example of wax effects on photolysis, either enhancement or retardation of photolysis was reported for 14C-fenarimol (239) spread
on glass surface with extracted waxes from barley, rape, strawberry, and apple
leaves (Watkins 1987). On the bean leaf, more S-oxidation of carboxin (42)
proceeded as compared with that on a glass surface (Buchenauer 1975). Accelerated photolysis was also reported for 2,4-D (1) on Zea mays leaves (Venkatesh
and Harrison 1999).
When 14C-dieldrin (123) was applied to bean foliage with flavone, formation
of photodieldrin (124) was enhanced by a factor of 57 via triplet photosensitization (Ivie and Casida 1971b). Rotenone exhibited the strongest sensitization
and effect of 15 pesticides tested, but anthraquinone had an insignificant effect.
The chromanone moiety of rotenone common to flavone was found essential to
show its activity (Ivie and Casida 1970). Accelerated photodegradation in the
presence of flavone, triplet sensitizers, and extracts from plant was also reported
for pyridafenthion (150), naproanilide (41), and flutolanil (39), all of which
were resistant to photolysis by themselves (Tsao et al. 1989; Tsao and Eto
1990a, 1991). Chlorophyll and furanocoumarins are also possible sensitizers in
plants, affecting photodegradation profiles of pesticides (Dodge and Knox 1986;
Dixon and Wells 1987). Although a reactive species has not been identified in
most cases, either possible hydrogen abstraction from wax components by the
photoproduced radicals originating from the pesticide or formation of a covalent
bond with the unsaturated bond of waxes via a radical reaction was demonstrated by using model waxes such as cyclohexene and methyl oleate (Draper
and Casida 1985; Schwack 1988; Schynowski and Schwack 1996; Breithaupt
and Schwack 2000). Nutahara and Murai (1984) have reported the enhanced
photodegradation of many pesticides by oleic and linoleic acids, possibly via
similar mechanisms as above. Because the unsaturated alkyl chain of these compounds is known to undergo oxidation by reaction with 1O2, giving the corresponding monohydroperoxide (Nakajima and Hidaka 1993), this peroxide or its
photodegradates may alter photodegradation of pesticides.

Photodegradation of Pesticides

17

Formation of possible photosensitizers on foliage has also been reported. 6Methyl-5-hepten-2-one (6-MHO) has been detected in significant amounts on
foliage of orange, oak, and pine trees together with 4-oxopentanal, geranylacetone, and acetone, whose origin is considered to be unsaturated wax components
such as sesquiterpenes and triterpenes (Fruekilde et al. 1998). By using epicuticular waxes extracted from these leaves, oxidation by O3 produced these carbonyl
compounds, possibly acting as an efficient photosensitizer on foliage. Berenbaum and Larson (1988) have reported the formation of 1O2 (4 1012 1O2
molecules cm2 sec1) by illuminating intact leaves of wild parsnip and prickly
ash. The reaction of ascorbic acid in plant cell walls with O3 was also found to
generate 1O2, and a similar reaction was found for gluthathione, methionine, and
uric acid (Chameides 1989). Kanofsky and Sima (2000) utilized chemiluminescence at 1270 nm due to 1O2 to monitor its formation under O3 from illuminated
tissue fluids prepared by vacuum infiltration technique or intact leaf of white
stonecrop. Emission of 1O2 with consumption of O3 was clearly demonstrated,
and its retardation in the presence of ascorbate oxidase showed involvement of
ascorbic acid in this reaction. Because O3 is a very familiar component of air
over vegetation, the photochemical and/or chemical generation of 1O2 may play
a substantial role in the degradation of some pesticides on foliage.

IV. Factors Controlling Photolysis on Soil Surfaces


A. Soil Components
Soil is a variable mixture of minerals, organic matter, and water, capable of
supporting plant life on the earths surface (Manahan 1994). The solid fraction
of a typical productive soil is approximately 5% organic matter, originating
from plant debris in various stages of decay, and 95% inorganic matter. Soil
usually contains air spaces and generally has a loose texture. The majority of
inorganic components (>90%) are crystalline and noncrystalline amorphous
minerals including primary and secondary minerals; the former includes quartz
and micas and the latter phyllosilicates (clay minerals), allophanes, and metal
oxides. Quartz and micas are simple SiO2 minerals whereas clay minerals are
basically aluminosilicates. Therefore, their surface property and photoreactivity
may be estimated from those of silica gel as a surrogate surface. There are three
types of hydroxyl groups existing on silica gel with a different acidity: geminal
silanol (Si(OH)2), nongeminal (SiOH), and hydrated silanol (SiOHOH2), as
demonstrated by 19Si- and H-NMR and fluorescence analysis (Thomas 1993).
Clay minerals possess layered structures consisting of silica tetrahedral and alumina octahedral sheets with a ratio of 1 : 1 or 2 : 1 (Takagi and Shichi 2000).
Kaolinite is the typical clay in the former type and pylophyllite, smectite, and
vermiculite groups constitute the latter. The isomorphous substitution of central
atoms in tetrahedral and octahedral structures with another of a lower valency
resulted in a net negative charge for clay sheets and electrostatic force via countercations, making them loosely bound to each other (Caine et al. 1999). The
presence of interlayer space thus gives a sterically constrained reaction environ-

18

T. Katagi

ment for pesticide molecules when intercalated. Iron is one of the most abundant
transition metals in soil and is considered to play a large role in photoinduced
redox reactions generating active oxygen species such as OH. Sherman (1989)
demonstrated by Mossbauer spectroscopy that most Fe3+ and Fe2+ ions were
found to occupy the octahedral sites and that some ions might occur as an
interlayer species such as a Fe2+-aquo complex in the 2 : 1 clay minerals. Instead
of the Fe3+-aquo complex, Fe(OH)2(H2O)4 would condense to form ferric hydroxy polymers.
Humic substances account for 60%70% of soil organic matter, consisting
of a series of highly acidic, yellow to black, high molecular weight polyelectrolytes and characterized by their high content of oxygen-containing functional
groups such as COOH, phenolic, aliphatic, and enolic OH and C=O, together
with amino, heterocyclic amino, imino, and sulfhydryl groups (Stevenson 1976;
Ruggiero 1999). The higher total acidity of fulvic acid (FA, 10 mEq g1) than
humic acid (HA, 7 meq g1) can be accounted for by the higher content of a
COOH group in FA. The typical content of each functional group was reported
to be 3.68.2 mEq g1 (COOH), 3.03.9 mEq g1 (phenolic OH), 2.66.1 mEq
g1 (alcoholic OH), 2.72.9 meq g1 (quinoid and ketonic C=O), and 0.60.8
mEq g1 (OCH3) (Choudhry 1984b). Similar results have been confirmed by
Senesi et al. (1989) by using several soil humic substances and Suwannee River
soil. The clay surface is usually covered with these humic substances, probably
via formation of claymetalorganic complexes. Through adsorption study of
atrazine (185), the contribution of clay surface on adsorption was proposed to
be important when organic matter content was less than 6% (Stevenson 1976).
Based on their nature, interactions with HA and FA should be primarily considered in soil photolysis of pesticides. As an adsorption isotherm, Freundlich,
Langmuir, and RothmundKornfeld equations are well known and the shape of
the related isotherm would shed light on the adsorption mechanism. Proposed
mechanisms of interactions are ionic bonding via cation exchange, hydrogen
bonding, charge-transfer interaction with a quinoid moiety embedded in a humic
structure, cation-bridged ligand exchange with humic carboxyl moiety, covalent
binding, hydrophobic adsorption, and partitioning via dipoledipole and/or van
der Waals interaction (Stevenson 1976; Senesi and Testini 1984; Senesi and
Miano 1995).
Another unique feature characteristic of soil humic substances is the presence
of stable radical species detected by ESR. Senesi and Schnitzer (1977) have
reported the ESR signal at g = 2.00322.0050 without a hyperfine splitting for
FA whose intensity increased with either chemical reduction under more acidic
conditions or UV irradiation. They proposed semiquinones or its ions as the
most likely partial structure of stable radical species. Further investigation on
various soil HAs and FAs has shown the presence of two types of ESR signals
originating from a quinhydron-type structure and a phenoxide ion (Choudhry
1981). These stable radicals would be involved in photoinduced transformation
of pesticides as well as formation of active oxygen species, especially when a
transition metal or its oxide coexists (Ruggiero 1999).

Photodegradation of Pesticides

19

B. Environmental Factors Affecting Soil Properties


Many soil photolysis studies have not strictly controlled and monitored basic
soil properties, especially such as moisture, until the recent development by the
Chibs group (Misra et al. 1997; Frank et al. 2002). As stated later, the penetration depth of light into soil is known to be very limited, approximately to the
top 0.5 cm of soil surface, and so its influence is considered to be very different
from that in the bulk soil (Fig. 6). In this region, solid, solution, and vapor
phases are all in close proximity to the soilatmosphere interface under sunlight
irradiation. Miller et al. (1989) concisely reviewed the effect of sunlight on
soil properties. They introduced a simulation on a typical diurnal variation of
temperature near the soil surface (02 cm) where the temperature increased up
to near 40 C at midday with a difference of about 20 during a day. A temperature gradient even at this shallow depth can be suspected and, in fact, the measured temperature at the irradiated surface, interior, and bottom of a soil thin
layer (0.51 mm thickness) attached to a cooling bath was reported to be 31,
29.9, and 25.6 C, respectively (Moore et al. 1989). This diurnal fluctuation of
temperature results in variation of soil moisture, that is, drying of soil during
daytime and rewetting during nighttime. These variations together with sunlight
irradiation would cause some effects on microbial activity, but its details are
not known for shallow soil. Recently, Reichman et al. (2000a,b) developed a
one-dimensional nonisothermal model to examine the dynamic behaviour of a
surface-applied pesticide under outdoor conditions. By using the diurnal variation of meteorological data such as wind speed, air temperature and relative
humidity and sunlight irradiation, they have simulated changes of depth-dependent soil temperature and moisture.
C. Mass Transport in Soil
Diffusion of a pesticide molecule is considered to be basically described by
Ficks law; however, the heterogeneous character of soil results in a more com-

Fig. 6. Structure of soil surface.

20

T. Katagi

plex description of a diffusion constant (D) than that in water (Graham-Bryce


1969). Do is the diffusion coefficient in free solution and is the tortuosity
factor:
D = Do v/(Kd d + v)
where v is the volumetric water content, Kd is the linear partition coefficient,
and d is the bulk density of soil. The tortuosity factor is a diffusion ratio of
pathlength in soil to that in aqueous solution (Scott and Phillips 1973). GrahamBryce (1969) has developed diffusion cylinder methods to determine the diffusion constants of pesticides. Information on the diffusion constant of a pesticide
in soil is very limited, but it seems to range from 0.05 to 50 mm2 d1 (Table 6
[see table on page 92]). When pesticide molecules are homogeneously distributed in uniform soil, the mean movement after time t is given by (2Dt)1/2. Therefore, it would take about 2.4 hr and 12.5 d for parathion (135) and trifluralin
(232) to move through the 1-mm-thick soil thin layer. The foregoing consideration shows the importance of mass movement in photodegradation on soil surfaces, but many other factors would operate in the field. Walker and Crawford
(1970) have demonstrated that diffusion constant was inversely proportional to
soil adsorption coefficient. Smaller diffusion with a higher water content was
reported for dinitroaniline herbicides (Jacques and Harvey 1979), whereas the
opposite relationship was observed for triazines (Scott and Phillips 1972). Ehlers
et al. (1969a,b) have reported the contribution to diffusion from vapor- and
nonvapor-phase mechanisms. In contrast to soil, several investigators have reported diffusion of pesticides in plant waxes and intact cuticles using the diffusion cell method (Schonherr and Riederer 1989; Lendzian and Kerstiens 1991;
Bauer and Schonherr 1992; Schreiber and Schonherr 1993), but their diffusion
seemed much slower than in soil.
Balmer et al. (2000) conducted the photolysis of trifluralin (232) and p-nitroanisole on kaolinite thin layers under constant temperature and humidity using
UV light at 300800 nm. By kinetic analysis, assuming first-order direct photolysis and diffusion following Ficks law, greater contribution of diffusion was
demonstrated for (232) than p-nitroanisole. For niclosamide (40), photodegradation in/on air-dried soil was greatly reduced in proportion to thickness of soil
thin layer, whereas a very slight increase of photolytic half-life was observed
for the moisture-maintained soil (Frank et al. 2002). Because the photic depth
of soil is usually less than 1 mm, as described in Section IV.D, migration of
(40) to the photic zone with the aid of soil moisture was considered most likely
to account for the observation. The importance of vapor-phase transport in airdried sandy loam soil was briefly examined for aryl ketones undergoing Norrish
type II photoelimination (Kieatiwong and Miller 1992). The existence of surfactant seemed to affect the migration of pesticides in soil. Gong et al. (2001)
showed that the faster photodegradation of atrazine (185) in the soil thin layer
with sodium dodecylbenzene sulfonate and proposed that solubilization of (185)
results in greater availability for photodegradation. A similar effect by hexade-

Photodegradation of Pesticides

21

cane in soil photolysis was reported for 2,3,7,8-TCDD (129) (Kieatiwong et al.
1990).
In addition, other factors such as the depth of the water table may have some
influence on pesticide movement. The upward movement of chlorsulfuron (96)
and triasulfuron (100) in a packed soil column was clearly demonstrated when
addition of water to soil surface and drying in a growth chamber were cycled
or the bottom of the column was dipped into water (Mahnken and Weber 1988).
Capillary movement of pesticide, similar to the latter case, was also reported
for norflurazon (214) and found to contribute to volatilization loss (Hubbs and
Lavy 1990). The effect of this type of upward movement on photolysis has been
confirmed for 14C-napropamide (47) in soil under sunlight at the different water
table levels (Donaldson and Miller 1996).
D. Photic Depth in Soil
Soils are highly heterogeneous and unmixed as compared with solution, and
illumination produces a light field difficult to define accurately (Wolfe et al.
1990; Senesi and Loffredo 1997). The depth of light penetration (photic depth)
in soil cannot be precisely defined but has been estimated by direct measurement
of transmitted fraction of light or using probe molecules selectively undergoing
direct or indirect photolysis. Transmission of xenon arc light through soil thin
layers with 0.17-mm thickness was measured and UV light was found to be
more than 90% attenuated (Herbert and Miller 1990). Frank et al. (2002) examined transmittance of UV light by varying the soil layer thickness from 0.5 to 4
mm. Even a 0.5-mm-thick soil layer was enough to block about 95% of the
incident light, but very slight light penetrated soil depths of 1.5 mm or greater.
There was no significant difference of transmission at three wavelengths (280,
365, and 440 nm). When soil thickness was less than 1.5 mm, more transmission
of light by a factor of 45 was observed for dry soil as compared with moist
soil, which may be accounted for by the difference in soil packing. As another
approach, Hebert and Miller (1990) utilized disulfoton (163) and flumetralin
(233) as chemical probes to estimate the photic depth in soil. Flumetralin (233)
absorbing light at wavelengths of 300500 nm undergoes direct photolysis,
whereas (163) has no UV absorption at >290 nm and only indirect photolysis
takes place, that is, oxidation to the corresponding sulfoxide via reaction with
1
O2. By using these pesticides as probes, mean photic depths were estimated to
be 0.23 and 0.28 mm for direct and indirect photolysis, respectively, in the
laboratory, with larger values obtained for the outdoor study, 0.32 and 0.62 mm,
respectively. The larger values for either indirect photolysis or outdoor study
may show contribution from diffusion of 1O2 to a deeper region of soil than
expected for light penetration and that from convective and evapotranspirative
transport of pesticide by thermal heating at the soilair interface. It is unclear
if light reduction originates from bulk attenuation or as an inner filter phenomenon by soil. If an adsorbing substrate is relatively porous and highly colored, it
would be possible for an adsorbed pesticide to be filtered from the incident

22

T. Katagi

light. Yokley et al. (1986) investigated photodegradation of pyrene and benzo on average),
[a]pyrene on silica gel, alumina, controlled pore glass (100 A
graphite, and various coal ashes by using a xenon arc lamp and reported the
importance of both adsorption to pores and UV screening. The UV screening
effect by soil was also reported for photodegradation of polyaromatics where
slower degradation was observed for soil with lower reflectance (Moore et al.
1989). Similar reduction of light has also been reported for sediments suspended
in aqueous solutions of pesticides (Miller and Zepp 1979; Oliver et al. 1979;
Zepp and Schlotzhauer 1981).
E. Effects of Soil Properties on Photolysis
Konstantinou et al. (2001) have shown the possible involvement of photosensitization based on faster photodegradation of herbicides in soil with higher organic
matter. In contrast, soil organic matter reduced the photodegradation of chlorimuron ethyl (102) (Choudhury and Dureja 1997a), triasulfuron (100) and thifensulfuron-methyl (105) (Albanis et al. 2002), and fipronil (220) (Bobe et al.
1998b), indicating involvement of either quenching or a shielding effect. For
niclosamide (40), modification of soil by addition of 3% HA was found to
reduce photodegradation (Graebing et al. 2002). An insignificant effect of organic matter on photolysis was observed for mecoprop (4) but its retardation by
soil amendment with 10% peat was detected under dry conditions, also implying
the quenching effect (Romero et al. 1998). Clay is the other important soil
component and may affect the photodegradation profiles of pesticides. Sukul
and Spiteller (2001) reported the linear relationship between photolytic half-life
of metalaxyl (37) and clay percentage in soil. Because (37) does not have any
significant UV absorption at >290 nm, photolysis was considered via indirect
photolysis, and the lower rate in soil with a higher clay percentage was likely
to originate from more intercalation of (37) into the intralattice structure of clay
where the incident light was screened.
Surface pH was an important factor, and such a pH effect would be operative
for acid-labile pesticides such as sulfonylurea herbicides (Schroeder 1997). One
of the other important factors to control photolysis would be soil moisture content. Faster photodegradation was detected in moistened than air-dried soils for
alachlor (34) (Fang 1977), imidazolinone herbicides (Curran et al. 1992), and
florasulam (48) (Krieger et al. 2000), and photoinduced hydrolysis and/or migration of pesticides to the photic zone of the soil thin layer might operate in these
cases. Enhancement of photolysis with soil moisture was observed for mecoprop
(4), but at higher moisture the photodegradation rates decreased (Romero et al.
1998). The significant decrease of photolysis rates in moistened soil was reported for fenpropathrin (24), which originated mainly from the increase of
surface pH on soil with moisture (Katagi 1993b). As clearly seen from these
examples, there are different mechanisms affecting photolysis with change of
soil moisture, and thus it is not easy to estimate its effect a priori.

Photodegradation of Pesticides

23

F. Photophysical and Photochemical Processes of Soil Components


Soil is a heterogeneous system where clay minerals are coated with humic substances and metal (hydr)oxides and the various photochemical reactions such as
photosensitization, quenching, and photoinduced reaction proceed (Fig. 7).
Humic Substances and Intact Soils There are many excellent reviews on photoprocesses of soil and aquatic humic substances (Zepp et al. 1981; Choudhry
1984a; Zepp 1988, 1991; Hoigne et al. 1989; Cooper 1989; Frimmel 1994).
UV-vis spectra of various HA and FA exhibited mostly featureless absorption
with their tail extending up to 500600 nm. The primary photoprocess is excitation to the short-lived (1 nsec) singlet states (Fig. 8). These excited states can
react with a pesticide molecule via a diffusional or static bimolecular process,
and the latter would play a greater role on soil surfaces. Zepp et al. (1985)
estimated the triplet energy of most humic substances to be around 60 kcal mol1
(250 kJ mole1) by using photoisomerization of 1,3-pentadiene as an index. It
depends on the energy level of triplet states of humic substances as compared
with that of a pesticide molecule which process predominates, sensitization or
quenching. Many pesticides are known to undergo photosensitized degradation
by humic substances (Jensen-Korte et al. 1987), whereas aqueous photodegradation of imazapyr (229) was slowed in the presence of HA due to the UV screening effect (Elazzouzi et al. 1999). In the photoinduced E/Z isomerization of four
alkyl cinnamates having a different association ability toward Fluka HA, only
2-fold increase in the rate of isomerization was observed for the cinnamate

Fig. 7. Photoreactions on soil surface: (A) sensitization, (B) quenching, (C) photoinduced
reaction, (D) adsorption/desorption, (E) spectral change, (F) inner-filter effect. P: pesticide; D: degradate(s); HS: humic substances; M: metal (oxides: hydroxides etc.).

24

T. Katagi

Fig. 8. Photophysical and photochemical processes of humic substances. HS: humic substances; RH: substrate.

having a 40-fold-higher affinity to HA (Van Noort et al. 1988), indicating that


the associated form was hardly available for triplet energy transfer.
The presence of reactive excited triplet states has been suggested for humic
substances through investigation of photoinduced degradation of cumene and
benzene (Kotzias et al. 1986). Because diffusion of pesticide molecules is considered to be restricted on soil surfaces, the possible reactions of a pesticide
molecule with an excited triplet of humic substances are transfer of electron or
hydrogen instead of direct energy transfer. These processes have been demonstrated for aqueous humic substances in photodegradation of the several methyl
and methoxy phenols (Canonica et al. 1995; Aguer and Richard 1996a). They
proposed an aromatic ketone moiety as a reactive site model in humic substances that underwent electron or hydrogen transfer via its excited n-* state.
Their contribution as well as energy transfer has been examined by using urea
herbicides including fenuron (51) and monuron (52) (Aguer and Richard 1996b;
Richard et al. 1997; Aguer and Richard 1999; Gerecke et al. 2001). Aguer et al.
(2001) showed the higher reactivity of HA fractions with a smaller molecular
size in photolysis of (51). They also demonstrated that soil FA having a smaller
specific absorption coefficient shows a higher activity, but that no meaningful
correlation is detected for HAs (Aguer et al. 2002). The weak photoinductive
efficiency of highly absorbing humic substances implied that the majority of
absorbing components had no photoinductive effect. There may be various small
organic molecules existing except for humic substances, and therefore they may
be involved in photolytic processes on soil. Kieber and Blough (1990) showed
the possible photoinduced formation of various carbon-centered radicals that
may react with pesticides under some photolytic conditions on soil surfaces.
Humic substances are known to react with O2 via energy or electron transfer
process to generate the very reactive 1O2, OH, superoxide anion (O2), and

Photodegradation of Pesticides

25

peroxide radical species (see Fig. 7). The contribution of each process was
roughly estimated for dissolved organic matter on an excited singlet state basis
(Zepp 1991). Thermal deactivation was a dominant process (97%99%), and
only 1%3% of energy would be transferred to an excited triplet state, most of
which was involved in formation of 1O2. In an aqueous solution of HA, Takahashi et al. (1988) confirmed the formation of OH, O2H, and 1O2 in the presence of O2 by ESR measurements using spin-trapping reagents. Nanosecond
laser spectroscopic methods were applied, and two short-lived transients from
four illuminated natural waters were estimated to be an excited triplet state of a
quinoid structure of humic substances and hydrated electron (eaq) (Fischer et al.
1985, 1987; Power et al. 1987; Zepp et al. 1987). Photoinduced generation of
eaq from natural waters also has been confirmed by its conversion to OH with
N2O (Thomas-Smith and Blough 2001), and the steady-state concentration of
eaq was estimated to be approximately 1.1 1017 M (Breugem et al. 1986).
Because eaq quickly reacts with water molecules to give O2 and O2H if generated (Hoigne et al. 1989), its importance would be greatly lessened, especially
on soil surfaces.
The steady-state concentration of 1O2 in natural water has been estimated to
be 1014 1012 M (Zepp et al. 1977; Wolfe et al. 1981; Haag and Hoigne 1986).
The quantum yield of the photoinduced generation of 1O2 from aqueous solutions of soil HA and FA was estimated by both ESR and chemical analysis
using furoin (1,2-dimethyl-2-hydroxyethanone) to be 0.395.5 103 (Aguer et
al. 1997). On soil surfaces, Gohre and Miller (1983) first demonstrated photoinduced formation of 1O2 by using tetramethylethylene and 2,5-dimethylfuran as
chemical probes. Gohre and Miller (1985) demonstrated that nontransition metal
oxide powders such as silica gel, aluminum oxide, and magnesium oxide catalyze the formation of 1O2. They proposed the reaction of an exciton bound to a
defect on a solid surface with adsorbed oxygen via transfer of electronic energy,
and no correlation between 1O2 generation and the content of soil organic matter
was reported (Gohre et al. 1986). The existence of 1O2 was supported by observation of chemiluminescence at 615650 nm possibly due to 1O2 dimoles, but
recently a direct spectrophotometric method for the very weak chemiluminescence at 1270 nm (1O2 3O2, spin-forbidden process) has been developed
(Rodgers 1987; Kanofsky 2000).
The reaction types of photosensitized oxidation via 1O2 can be classified into
formation of endo-peroxide, ene reaction giving allyl hydroperoxide, and formation of 1,2-dioxetane (Foote 1968a,b; Wilkinson and Brummer 1981). The
thioether-containing pesticides on soil were found to be transformed to the corresponding sulfoxide by sunlight exposure most likely via reaction with 1O2
(Gohre and Miller 1986). The other example was photoinduced degradation of
bioresmethrin, 1R-trans isomer of (15), whose degradation was clearly reduced
in the presence of -carotene as an efficient 1O2 quencher (Clements and Wells
1992). The enhanced degradation of 2-dimethylamino-5,6-dimethylpyrimidin-4ol in D2O inhibited by sodium azide also showed involvement of 1O2 in its
degradation, possibly attacking the 5,6-double bond (Dixon and Wells 1983).

26

T. Katagi

Under exposure to sunlight, OH is photochemically generated via (1) reaction of humic substances in the excited triplet with water, (2) dissociation of
nitrate ion followed by protonation, and (3) degradation of H2O2 formed by
reaction of eaq with water (Hoigne et al. 1989; Zepp 1991); in addition, coexistence of metal cations such as ferric ion with humic substances was considered
to play a great role in generating OH via (4) the photo-Fenton reaction. The
second mechanism is considered to be predominant in the aquatic environment,
but investigation has demonstrated that about half of OH is generated via
photo-Fenton reaction, and an O2-independent pathway also exists (Vaugham
and Blough 1998). Among these four mechanisms, the first and last procedures
are considered to be more important to soil photodegradation. Photoinduced
formation of OH was confirmed in photolysis of 2,6-dimethoxyphenol with soil
extracts by ESR (Suflita et al. 1981). Many investigations on OH-generating
ability of metalhumate complex through the Fenton or HaberWeiss mechanism have been reported. Paciolla et al. (1999) demonstrated the intrinsic ability
of Fe3+ or Cu2+ HA complexes to generate OH from H2O2 in darkness. Mossbauer spectroscopy of this complex showed that about 5% of iron ion was present as Fe2+, indicating involvement of the above mechanisms.
Fukushima and Tatsumi (1999) studied the photocatalytic activity of the
Fe3+ soil humate complex giving OH and H2O2 at >370 nm. Photoinduced oneelectron transfer from HA to O2 followed by protonation was considered to yield
O2H, which was then disproportioned into H2O2 and O2. They demonstrated
that the majority of Fe3+ was complexed with the HA fraction in a higher molecular weight and that the substrates were incorporated into the molecular skeleton
of HA through photolysis (Fukushima et al. 2000, 2001; Fukushima and Tatsumi 2001). Similar photo-Fenton reactions can be considered for small organic
molecules (Zepp et al. 1992; Balmer and Sulzberger 1999). Because soil components such as humic substances and silicate surface can trap a peroxy radical
(Pohlman and Mill 1983), contribution from the aforementioned processes may
be lessened on soil surfaces.
As described, H2O2 is an another important reactive oxygen species in the
environment whose generated humic substances are known to be deeply involved, and its steady-state concentration in an aquatic environment is estimated
to be 105 107 M (Draper and Crosby 1983; Cooper 1989). H2O2 on soil surfaces would be degraded by various interactions with soil components or react
with a pesticide molecule under sunlight irradiation. Petigara et al. (2002) investigated the degradation processes of H2O2 in soils and revealed the possible
involvement of three pathways: a metal-catalyzed HarberWeiss reaction, twoelectron process achieved by catalase, and a peroxidic-type reaction. OH was
fluorometrically determined using spin-trapping reagent and fluorescamine
(Vaugham and Blough 1998). Retarded loss of H2O2 by autoclaving or addition
of sodium azide or formaldehyde showed that 65%75% of loss was due to
abiotic processes on soil particles. Among minerals, goethite was found more
reactive in generation of OH than hematite. The photoinduced reaction of H2O2
with pesticide was not reported on soil surfaces, but aqueous photolysis of the

Photodegradation of Pesticides

27

several pesticides in the presence of H2O2 clearly showed involvement of photogenerated OH as evidenced by product analysis (Draper and Crosby 1981,
1984).
Clays The flat organosilicate sheets with reactive edges give a unique reaction
environment on clay. Furthermore, a geometric constraint by pore or sheet structures in clay may alter the photochemistry of pesticides (Thomas 1993). A clay
surface may exhibit a cage effect for radicals generated via photoinduced homolytic cleavage of a bond similarly observed for benzyl derivatives on silica gel
(Avnir et al. 1981). As one of the basic properties on clay surfaces, it should be
noted that the sites possessing a high Brnsted acidity may show catalysis on
some pesticides (Caine et al. 1999). Using these properties, Takagi and Shichi
(2000) reviewed organic photochemistry in/on the clay surface. The efficient
transfer of electron or excited energy occurs between molecules adsorbed on
clay surfaces, and ferric ion as an impurity in clay may act as an efficient
quencher. Spectral shift by adsorption, steric constraint, and hydrogen bonding
are known to result in the different profiles for photoinduced isomerization of
styrene derivatives, Norrish type I and II reactions of aromatic ketones, and
photo-Fries rearrangement of carbamates.
One-electron transfer from an adsorbed chemical to clay is known for thianthrene on laponite, as evidenced by ESR (Mao and Thomas 1993). A similar
photoinduced electron transfer to laponite, silica gel, and silica-alumina was
reported for pyrene (Liu et al. 1994). Extent of a cation radical formation was
found to gradually decrease at above 340 nm, indicating the existence of variable active sites having a different capacity of accepting electrons. There is also
evidence that ferric ion included in montmorillonite clay as an impurity acts as
an electron mediator between excited 10-methylacridinium hexafluorophosphate
and iodide as an acceptor (Theng et al. 1997). The other type of electron transfer
is directed from clay surface to adsorbed species such as O2 and transition metal
ions in clay is considered to be deeply involved, as already described.
Various kinds of energy transfer on clay have been investigated. Detailed
analysis of difference IR spectra implied the possible interaction of bioresmethrin, 1R-trans isomer of (15), with methyl green (MG) on clay surfaces, and
photostabilization of bioresmethrin by MG could be accounted for by an efficient energy transfer (Margulies et al. 1985, 1987, 1993). Margulies et al. (1988)
have reported the red shift of absorption spectrum when NMH (242) was adsorbed on montmorillonite. Its photostablity was improved by coadsorption of
thioflavin (TFT) or MG and its extent was greater in the former dye. The UVvis absorption maximum of TFT (410 nm) was closer to that of (242) compared
to MG (630 nm), implying more efficient energy transfer from (242) to TFT.
The interaction of (242) with clay surface was considered to take place at the
cyclic enamine moiety of (242) as estimated by difference IR spectrum. Furthermore, the direct energy transfer to clay surface was also reported for (242)
(Rosen and Margulies 1991). Less photostabilization of (242) adsorbed to nontronite than hectorite and montmorillonite could be accounted for by the differ-

28

T. Katagi

ent contents of Fe3+ (29.46%, 0.26%, and 3.72% as Fe2O3, respectively), which
acted as a quencher via a charge-transfer mechanism. Efficient overlap of the
emission with the absorption spectrum is indispensable for this energy transfer,
which has been well established for norflurazon (214) coadsorbed with TFT on
montmorillonite (Undabeytia et al. 2000). They also proposed the photoinduced
generation of OH by montmorillonite accelerating its degradation. The involvement of reactive oxygen species in photodegradation of pesticides on clay surfaces has been reported for tolclofos-methyl (142), esfenvalerate (28), and PBacid (243) (Katagi 1990, 1991, 1992). A simple organic cation such as
benzyltrimethyl-ammonium has been demonstrated to modify adsorption of
alachlor (34) to a clay surface, leading to its protection against photolysis (ElNahhal et al. 1999). The IR spectral shift of carbonyl and C(aromatic)-N bonds
to lower and higher wave numbers, respectively, indicated the possible interaction of carbonyl oxygen with an exchangeable countercation on the clay surface
through a water bridge (Nir et al. 2000). The red shift of the UV absorption
spectrum (-*) of phenyltrimethylammonium was observed when coadsorbed
to montmorillonite, indicating the interaction between the phenyl rings of (34)
and the organic cation.
In addition, clay surfaces also give a steric constraint on an adsorbed pesticide molecule (Margulies et al. 1993; El-Nahhal et al. 2001). Photostabilization
of trifluralin (232) was realized by adsorption to montmorillonite, but the addition of TFT did not give further improvement (Margulies et al. 1992). Difference IR spectra clearly demonstrated interaction of one nitro group of (232)
with the clay interface. It is known that (232) undergoes photo-induced cyclization between nitro nitrogen and the C1 carbon of the N-propyl moiety to give
benzimidazole derivatives (see Fig. 2, reaction 10b). Because the intramolecular
cyclization requires a conformational change of the N-propyl moiety, which is
restricted by adsorption to the clay, the observed photostabilization is most
likely to be accounted for by imposition of a steric constraint on (232).
Metal Oxides and Hydroxides Transition metals such as iron and manganese
exist in the environment as their oxides, hydroxides, and sulfides or impurities
in clay being substituted with aluminum ions. These compounds can become
quenchers of an excited energy and catalytic sites by acting as an electron mediator. It has been reported on silica gel that dye-sensitized photooxidation of
bromacil (198) by 1O2 is significantly reduced by addition of FeO, Fe2O3, Fe3O4,
or FeO(OH), which can be accounted for by an electron transfer from the excited dyes or energy transfer from 1O2 (Riter et al. 1990). In addition, as represented by titanium dioxide, some materials also can act as semiconductors where
the photoexcited valence band electron reduces an organic chemical and the
accompaning hole oxidizes (Balkaya 2003). Irradiation at >340 nm accelerated
degradation of (185) in the presence of typical semiconductors TiO2, ZnO, and
WO3 to give N-dealkylated derivatives possibly via oxidation by OH, whereas
-Fe2O3 and Al2O3 did not show any photocatalytic action (Pelizzetti et al.
1993). Further examination was conducted by Lackhoff and Niessner (2002) to

Photodegradation of Pesticides

29

estimate such a photocatalytic activity for environmental particles. Both TiO2


and ZnO exhibited a significant acceleration of degradation by a factor of 30
300, but much less activity was detected only for the other oxides such as
SrTiO3, Fe2O3, and FeTiO3. Fine particles from natural sand, soot, and ash did
not show any meaningful acceleration compared with direct photolysis, which
would be due to less occurrence of photocatalytic Ti (<1%) in these materials.
Only very weak photocatalytic activity was observed for three soils when (185)
was irradiated at >340 nm in aqueous suspension (Pelizzetti et al. 1993).
Similar trends of photocatalysis were briefly reported for dicarboximide fungicides (Hustert and Moza 1997) and azo dyes (Hustert and Moza 1994). In
contrast, the direct reaction with a positive hole was proposed for photolysis of
2,6-dimethylphenol in aqueous suspension of goethite (-FeOOH) (Mazellier
and Bolte 2000). Not only the surfaces of these metal (hydro)oxides but also
species dissolved from the surface are considered to show catalytic acitivity in
generating active oxygen species. Voelker et al. (1997) demonstrated this in the
mixture of lepidocrocite (-FeOOH) and FA, and Fenton-like degradation of
H2O2 has also been reported for goethite (Chen and Watts 2000). Many iron
species are considered to exist in an aqueous phase after dissolution, and
Fe(OH)2+ species have been found most efficient in generating OH under irradiation (Feng and Nansheng 2000). When H2O2 is enzymatically or abiotically
formed in soil, OH is likely to be produced via Fenton-like reaction as demonstrated by ESR (Huling et al. 1998). These reactions may occur in real soil and
contribute, at least in part, to photodegradation of pesticides. In contrast to iron
species, manganese oxides seem to show much less photocatalytic activity (Bertino and Zepp 1991; Lee and Huang 1995).

V. Atmospheric Oxygen Species


Reaction with oxygen species in air at a solidair interface is considered to be
another important degradation pathway for pesticides deposited on soil and plant
surfaces. In the troposphere, O2 is the most abundant reactive species from its
biradical structure in the ground state and is involved in either autooxidation or
coupling with radical species to give the corresponding peroxy species. The
other important species are O3 and OH because of their high reactivity (Crosby
1983; Marcheterre et al. 1988). In fact, their possible contribution to degradation
of pesticides has been demonstrated on plant foliage (Spear et al. 1978; Ophoff
et al. 1999) and soil surfaces (Spencer et al. 1980; Kromer et al. 1999). 1O2 is
also the other candidate, but its contribution may be limited to its specific reactivity toward a thiol moiety and C=C double bonds (Wilkinson and Brummer
1981; Larson and Marley 1999). The formation and decline processes of these
reactive species in air are known to be very complex (Prinn 1994). O3 in the
troposphere originates from that in the stratosphere formed by direct photolysis
of O2 or reaction between O2 and singlet oxygen atom O(1D) and dissipates via
UV photolysis to O(1D) or reaction with NO. OH is mainly generated via reaction between O(1D) and H2O and deactivated by reaction with air pollutants

30

T. Katagi

such as CH4, SO2, and CO. The lifetime of O3 and OH in air is estimated to be
3 105 sec and 2.7 sec, respectively.
Concentrations of O3 and OH are significantly low and known to vary with
many factors such as climate, geographical features, vegetation, and pollution
by human activities. The monthly mean value of troposheric O3 was 3550
ppb in the western North Atlantic (Oltmans and Levy 1992). The yearly mean
concentration at 13 stations in the United States was about 30 ppb with a seasonal variation, the maximum value (8090 ppb) being detected in spring (Singh
et al. 1978). The diurnal variation was also reported, and the maximum concentration was observed in the morning due to low photochemical destruction (Oltmans 1981). Logan (1999) found that rural sites downwind of urban areas might
have particularly high values of O3 in the middle of the day; 4070 ppb in
summer and 2030 ppb in winter. The steady-state concentration of OH was
estimated to be 56 105 molecules cm3 (Crutzen 1982). Direct measurements
have been conducted by using laser long-path absorption or laser-induced fluorescence method based on the X2 ( = 0) A2+ ( = 0) transition of OH.
The former method gave the OH concentration of 4.3 (0.4) 106 molecules
cm3 in a sunny period but less in the nighttime (<1.5 106 molecules cm3)
(Dorn et al. 1995). The latter method showed similar values (3.46 106 molecules cm3) in May and June (Holland et al. 1995). As laser irradiation also
produces OH from the coexisting O3 as an artifact, ion-assisted mass spectrometry using 34SO2 has been applied (Crosley 1995). Estimated values in Colorado
were 24 106 molecules cm3 smaller than expected from a photochemical
model, which could be accounted for by the presence of a quencher such as
isoprene (2 ppb) in air. In other places, 110 106 molecules cm3 was correctly
estimated.
O3 and gaseous pollutants in air are considered to be sorbed to plants and
soil surfaces, and such deposition is one of the important factors controlling
their air concentrations. In the case of plant foliage, not only the direct interaction of O3 and OH via sorption or reaction with epicuticular waxes but also
reaction with epidermal and mesophyll cells and substomatal cavities after passage through stomata is considered to play a role (Runeckles 1992; Cape 1997).
The soil surface also acted as a sink of O3 (Turner et al. 1974; Fontan et al.
1992; Cieslik and Labatut 1997). The uptake of O3 by plants was considered to
be mainly through stomata, and its flux over the vegetation was monitored (Eastman et al. 1981). As a result, the deposition velocity over maize (max., 0.84 cm
sec1) was found to be about twice as large as that over grass, suggesting that
the size of the stomatal aperture may be the predominant mechanism in O3
uptake. Wet and dry O3 deposition via stomatal or nonstomatal mechanisms
have been studied in conjunction with aerodynamics over vegetation (Fowler
et al. 2001; Zhang et al. 2002). Kerstiens and Lendzion (1989) examined nonstomatal deposition of O3 using isolated cuticle from various plant leaves and
demonstrated that its permeance via leaf is highly dependent on plant species
and that either dust or hairs on leaf surface provides more degradation sites for
O3. Based on these considerations, atmospheric conditions, especially for the

Photodegradation of Pesticides

31

concentration of O3 and OH, should also be taken into account when photodegradation pathways of pesticides on plant and soil surfaces are examined.

VI. Experimental Design and Kinetic Analysis


A. Light Source
Many kinds of conventional light sources with different spectral irradiance
(Gould 1989a) have been utilized especially for photolysis studies of pesticides
(Guth 1981; Roof 1982; Miller and Zepp 1983; Parlar 1990). The spectral comparison of artificial light with natural sunlight indicates great differences (Fig.
9) (Hirt et al. 1960). The most suitable light source is a xenon arc lamp; as it
emits at <290 nm, an appropriate UV filter should be used. The glass and solution filters commonly used in photolysis studies are each characterized by transmission of light at a specific wavelength range (Gould 1989a), and glass has
been preferred because of its simplicity. The most popular filter is Pyrex (or
Duran) glass and its thickness (4 mm) is also important for effective cutoff of
the undesired shorter wavelengths (Zepp 1982). Cellulose acetate sheet is sometimes unfavorable due to its solarization. If an appropriate glass and/or solution
filter is used, the wavelength dependency of photolysis can be conveniently
estimated for better understanding of the photolytic mechanism (Schwack and

Fig. 9. Spectral irradiance of typical light sources used in photolysis studies. Graph is
based on data from Hirt et al. (1960).

32

T. Katagi

Kopf 1993; Schwack et al. 1995c). Even if natural sunlight is used, it should be
kept in mind that a reaction vessel or the glass of the greenhouse can partially
absorb sunlight.
B. Photolysis Chambers
Ebing and Schuphan (1979) introduced a closed cultivating system made of
Pyrex or Duran glass (3-mm thickness) and equipped with volatile traps under
irradiation by fluorescent tubes to examine degradation profiles of dichlofluanid
(205) with a good 14C recovery (96%99%) in soilplant (spinach, potato, and
cress) systems. A volatilization chamber elaborately designed to simulate well
real environmental conditions was developed to evaluate the dissipation profiles
of methyl parathion (136) applied to French beans (Muller et al. 1995). The
chamber was connected with an air-conditioning unit to obtain the desired temperature, humidity, and flow rate of air that passes over the plants into volatile
traps, and the ceiling of the chamber was made of a special glass transmittimg
light from a metal-halogenide lamp to simulate exposure to natural sunlight. A
similar wind tunnel apparatus was successfully introduced to estimate volatilization and degradation profiles of fenpropimorph (227) on dwarf beans, sugar
beet, and radish (Ophoff et al. 1999). However, these apparati are not readily
available and their maintenance is difficult; therefore, photodegradation of pesticides on plants has been mostly examined with metabolism studies.
Many researchers have developed an apparatus for soil photolysis that can
be classified as a rotatory reaction vessel made of glass or thin-layer plate in a
chamber (Guth 1981; Choudhry and Webster 1985; Parlar 1990). The former
system was introduced to examine the mineralization of pesticides on a silica
gel surface (Parlar 1990), and its concept has been used to examine photolysis
of diuron (53) adsorbed on sand, clays, and iron oxide (Jirkovsky et al. 1997).
The latter system has been widely utilized, and its common design is illustrated
in Fig. 10, based on the literature (Klehr et al. 1983; Katagi 1990; Misra et al.
1997; Kromer et al. 1999; Balmer et al. 2000). The soil thin layer (thickness <2
mm) prepared on a plate or vessel made of glass or stainless steel is attached to
the controlled temperature water bath, and sometimes a thermocouple buried in
the soil is utilized to monitor and control the soil temperature. Light intensity
and its spectral distribution are monitored by a spectroradiometer. Because either diffusion of a pesticide molecule or photodegradation profiles in or on soil
is very sensitive to soil temperature and humidity, some investigators (Misra et
al. 1997; Kromer et al. 1999; Graebing et al. 2003) have also introduced a
relative humidity sensor to the chamber equipped with a water spray nozzle and
established a computerized system to automatically control humidity.
Good material balance is indispensable for precisely evaluating the photolytic
profiles of pesticides and thus the selection of appropriate traps is very important. Not only do volatile pesticides have a higher vapor pressure, most volatiles
consist mainly of CO2 and organic degradates with a small molecular size. Air
flow under reduced pressure conditions was obtained using a suitable pump to

Photodegradation of Pesticides

33

Fig. 10. Soil photolysis chamber. 1, soil thin layer (<2-mm thickness); 2, thermocouple;
3, thermometer; 4, photoprobe connected by glass fiber; 5, spectroradiometer; 6, glass
filter (Pyrex); 7, thermostat-controlled water bath; 8, circulator; 9, xenon arc lamp; 10,
power supply; 11, CO2-free humidified air; 12, traps.

avoid undesirable leakage from glass joints or seals instead of the pressurized
system. One of the traps is a liquid type, represented by aqueous alkaline solution and organic solvents and the other is a solid type, for example, ascarite,
charcoal, and porous polymers (Lewis 1976). CO2 is usually collected by using
0.10.5 M KOH or NaOH in a gas-washing bottle and quantified by precipitation as BaCO3. Monoethanolamine is also used as a trapping medium (Yamaoka
et al. 1988; Tanaka et al. 1991), and its efficiency can be improved by adding
methanol and a scintillation cocktail (Abbott et al. 1992). Ascarite has been
conveniently used to trap CO2 generated during photolysis of florasulam (48)
(Krieger et al. 2000). When CO is generated, it can be catalytically converted
to CO2 using a combustion furnace (CuO at 650 C) followed by trapping as
above (Tanaka et al. 1991) or chemically trapping as cuprous complex using an
acidic CuCl solution (Busch and Franklin 1979). Small organic volatiles are
usually collected by using nonvolatile ethylene glycol, and its low solubilizing
ability toward organic molecules can be improved, for example, by addition of
a small volume of xylene. Evaporation of an organic solvent usually causes
difficulty in its use as a trapping medium, especially for longer periods, but
cooling of the medium can improve its utility (Koshy et al. 1983). The cryogenic
technique using a dry iceacetone mixture at 78 C can also be effective
(Lewis 1976). As a solid trapping medium, acetone-washed polyurethane foam
(Kromer et al. 1999) and porous polymeric sorbent such as Amberlite XAD
resins, Chromosorb, and Tenax (Lewis 1976; Smith et al. 1995; Ophoff et al.

34

T. Katagi

1999) have been used effectively. If organic volatiles can undergo fast and specific chemical reaction in a trapping medium, it can become a useful trapping
method. Tanaka et al. (1991) used dimedone (5,5-dimethylcyclohexane-1,3dione) with a trace amount of pyridine to trap formaldehyde. For formic acid
and acetic acid, Yamaoka et al. (1988) derivatized these using p-bromophenacyl
bromide in the presence of 18-crown-6.
C. Kinetic Analysis
When the reaction environment of a pesticide varies with its movement and
adsorption, a simple first-order kinetics cannot be applied. In the foliar dissipation of pesticides, the decline curve sometimes follows a two- or three-phase
kinetics, each of which consists of a single exponential decline (first order).
These are considered to correspond to the compartment models where movement of a pesticide between each compartment can be neglected. Gunther (1969)
demonstrated the involvement of two dissipation processes for pesticides applied to citrus fruit and found that the first faster dissipation stems from surface
deposits in epicuticular waxes and the slower process from metabolism in the
rind. Similar approaches are utilized for many studies. Figure 11 shows the differences in decline profiles among simple first-order (single-phase), two- and
three-phase kinetics as examples. By considering a weight of vaporization loss
from plant surfaces in the early stage after application, Stamper et al. (1979)
analyzed crop residue data by ln-ln plots and statistically obtained the better
relationship; R = a*t3/2 (R, residue; t, days after application; a, constant). When
Ficks second law of diffusion is applied, the dissipation of a pesticide can be
approximated by y0 /(4Dt)3/2 (y0, initial deposit; D, diffusion constant), which
is in accordance with the statistical analysis. This information indicates the importance of controlling factors such as volatilization besides basic degradation
mechanisms, and such an approach has been undertaken to account for dissipation of pesticides from tea plants (Zongmao and Haibin 1997). The quantum
chemical parameters were also introduced to examine photodegradation of polychlorodibenzodioxins on laurel cherry (Chen et al. 2001), and such an approach
may be useful.
In the case of soil photolysis studies, the decline of a pesticide usually does
not follow the first-order kinetics and slows down with time, possibly due to
either adsorption to soil or movement out of a photic zone. Many investigators
have applied the single- to three-phase models, but in some cases the secondorder rate constant or Hoerl function [a * exp (b*t) * tc], where a, b, and c are
constants and t is an incubation period, has been proposed as a better relationship (Emmelin et al. 1993; Romero et al. 1998). The simpler equation based on
a meaningful physicochemical assumption is desirable in the usual analysis of
experimental data. Gustafson and Holden (1990) have focused on a spatial variability in factors affecting dissipation rate in soil such as microbial population,
light intensity, temperature, and soil properties. They assumed an infinite compartment model where each compartment with a different dissipation rate is

Photodegradation of Pesticides

35

distributed with some probability and introduced the concept that the rate constants follow the distribution. The simple equation, y = (1 + b*t)a, can be
derived with a half-life of (0.5(1/a) 1)/b (y, % of the applied dose; t, time; a, b,
constants); the dissipation curve is shown in Fig. 11. Data analysis using these
approaches helps to understand not only persistence of a pesticide but also the
dissipation mechanisms involved.

VII. Photodegradation of Pesticides in Model Systems


A. Soil Surface Models
Glass, silica gel, or clay have been utilized as simple models by considering
heterogeneity and variability of soil affecting photodegradation profiles of pesticides (Hulpke et al. 1983). However, the adsorptive nature of soil cannot be
taken into account on glass. This effect may be partly realized in silica gel or
clay, but the higher reactivity of their surface originating from many types of
silanol groups and adsorbed water would give different reaction environments.
Furthermore, more tightly packing may reduce the depth of light penetration
compared with soil. Although the contribution of soil humic substances acting
as a photosensitizer, quencher, or solubilizing medium is completely neglected
in either of these models, their easier handling makes them the first useful approach to evaluate soil photolysis of pesticides. Treatment of silica gel or clay

Fig. 11. Typical decline curves. When period and percent (%) of the applied dose are
t and y, respectively, each curve can be defined as follows: single-phase, y =
exp(a*t), a = 0.139; two-phase, y = a*exp(b*t) + c*exp(d*t), a = 0.6, b = 0.139,
c = 0.4, d = 0.0139; three-phase, y = a*exp(b*t) + c*exp(d*t) + e*exp(f *t), a = 0.4,
b = 0.139, c = 0.3, d = 0.0139, e = 0.3, f = 0.00693; Gustafson, y = (1 + b*t)a, a = 1.5,
b = 0.117.

36

T. Katagi

with humic substances can be alternatively considered, but such approaches are
very limited (Schafmeier et al. 1998). The most frequently reported system is
the aqueous soil suspension (Pelizzetti et al. 1990; Mansour et al. 1997; Hustert
et al. 1999), but the excess amount of water may alter the reaction environment
or unexpectedly increase the contribution either from direct or indirect photolysis in an aqueous phase containing dissolved humic substances. Furthermore,
the filter effect by suspended matters may reduce the rate of photolysis (Zepp
and Schlotzhauer 1981), and the adsorption to soil may be underestimated for
partially water-soluble pesticides. Furthermore, even if the soil thin layer with
1- to 2-mm thickness is used, diurnal variation of soil temperature and moisture
is difficult to determine (Miller et al. 1989; Reichman et al. 2000b). These
differences would change the extent of vaporization loss of pesticide from soil
surface and misstate the rate of photoinduced dissipation by neglecting transport
of pesticide or reactive oxygen species via diffusion along with the movement
of water or air in soil (Ehlers et al. 1969b; Hubbs and Lavy 1990; Donaldson
and Miller 1996; Balmer et al. 2000). The few approaches elaborately controlling these factors have been carefully developed (Misra et al. 1997; Kromer et
al. 1999; Frank et al. 2002).
B. Plant Surface Models
The simplest model is the glass surface, but the reaction environment is very
different from the cuticular surface. From this point of view, Peacock et al.
(1994) have briefly examined an inactive polytetrafluoroethylene plate. Considering the wax chemistry described in Section III.E, many researchers have conveniently examined photodegradation of a pesticide in simple organic solvents
as surrogates of waxes, but their significantly different fluidity and simple structures should be kept in mind together with easier photoaddition of a solvent
molecule. Schwack (1988, Schwack et al. 1994) used methyl oleate and 12hydroxystearate as more elaborate models of epicuticular wax. Photolysis of
2,4-D (1) has been reported to be enhanced on Zea mays leaves and thus a more
realistic model has been considered (Venkatesch and Harrison 1999). Photodegradation of carbamate and organophosphate pesticides was examined in thin
film of epicuticular waxes extracted from a variety of citrus fruits (Cabras et al.
1997b; Pirisi et al. 1998, 2001). Schuler et al. (1998) examined photodegradation of dibenzo-p-dioxins including (129) on glass coated with waxes of laurel
cherry leaves. The photolysis rate of fenarimol (239) was dependent on plant
species (Watkins 1987). Because plant cuticles can be isolated from intact leaves
by treatment with pectinase and cellulase, Schynowski and Schwack (1996) conducted photodegradation of parathion (135) on enzymatically isolated cuticles.
C. Photodegradation of Pesticides on Glass and Silica Gel Surfaces
Glass plates (Chukwudebe et al. 1989) and silica gel (Hirayama et al. 1998)
were utilized to examine the comparative photoreactivity of various pesticides.
Any significant relationship between the photolysis rate on glass at max = 300

Photodegradation of Pesticides

37

nm and the extinction coefficient of each pesticide at 295305 nm could not be


identified, but the slower degradation was observed for the thicker layer of
pesticide on glass (Chen et al. 1984). Photolysis rates were greatly reduced at
the concentration of 3.3 g cm2, corresponding to the thin-layer thickness of
. Since Kitchener (1946) reported that most of light passing through a
100 A
, this reduction was likely to
crystal is absorbed along the pathlength of 100 A
originate from light attenuation. Furthermore, the different profiles of UV absorption in the solid state from those in solution (Gab et al. 1975b; Leermakers
et al. 1966) may account, at least in part, for this insignificant relationship. In
contrast, Calumpang et al. (1984) found a good correlation between maximum
molar absorptivity in solution and photolysis rate for eight organophosphate
insecticides on silica gel. On these solid surfaces, the photoinduced homolytic
cleavage of a bond to produce radicals is considered to be one of the most
important degradation pathways. Johnston et al. (1984) studied photodegradation of azobis(isobutyronitrile) on dry silica gel by using UV light and demonstrated that some portion of radicals could escape from their original partners
by a translational motion. Avnir et al. (1981) also examined the reaction profiles
of radicals by using benzyl phenylacetate and dibenzyl ketone. By comparing
the amount of dibenzyl formed in potassium dodecanoate micelles via recombination of radicals, they concluded that radical pairs separated more easily on
silica gel surface by a translational motion. Furthermore, this trend was more
emphasized for dibenzyl ketone generating radicals via the triplet state compared with benzyl phenylacetate via the singlet state. Therefore, if the pesticide
molecule deposited on silica gel surface undergoes photoinduced homolytic
cleavage of some bond, the recombination process may be less important. The
photodegradation profiles of pesticides on glass and silica gel are summarized
in Table 7 (see table on page 94) from a survey of the existing literature.
Organochlorines Photoinduced dechlorination of pentachlorophenol (121)
gave the -radical and successively the phenoxide (Piccinini et al. 1998). The
photolytic half-life was dependent on the solvent used to prepare a thin film on
glass, and a residual amount of water was found to increase the photolysis rate.
Aldrin (122) on glass underwent epoxidation to dieldrin (123) followed by further isomerization to photodieldrin (124) (Rosen and Sutherland 1967). In the
case of endosulfan (125), the ring opening to form the diol derivative via release
of SO was a major pathway but without S-oxidation (Dureja and Mukerjee
1982). The other cyclodienes, including chlordane (128), were found to form
the corresponding cage structure derivatives (Benson et al. 1971). More epoxidation and hydroxylation of (128) on silica gel implied involvement of reactive
oxygen species on the surface formed by irradiation (Gab et al. 1975a). DDT
(130) on glass primarily underwent photoinduced homolytic cleavage of the
CCl bond at the trichloromethyl group to form DDD (132) by exposure to UV
light at 254 nm (Mosier et al. 1969). The radical mechanism was also supported
by comparative photolysis of eight 3H-diethoxy analogues of (130) on glass
(Coats et al. 1979). Relative photostability decreased by substituting the CCl3

38

T. Katagi

moiety with other groups as follows: CCl3 (130) >> CHCl2 (132) > CHCl2CH3
>> CCl(CH3)2. This order was in agreement with the stability of radical species
formed via CCl cleavage.
Organophosphorus Esters Photoinduced oxidation of the P=S moiety was observed for parathion (135) and its methyl analogue (136) on glass (El-Refai and
Hopkins 1966; Calumpang et al. 1984), together with formation of S-isomers
(Chukwudebe et al. 1989). Fenitrothion (138) on silica gel underwent oxidation
at the aryl methyl group to COOH and/or P=S moiety to the oxon together with
an ester cleavage giving the corresponding phenol (Ohkawa et al. 1974b). ODemethylation was observed for the rapid photodegradation of cyanophos (137)
on silica gel in addition to oxon formation and cleavage of the PO aryl bond
(Mikami et al. 1976). The photoinduced S-oxidation of the methylthio group to
sulfoxide was the primary reaction on glass for fenthion (143) (Cabras et al.
1997b). Hirahara et al. (2001) examined comparative photodegradation of (143)
using UV-A (320400 nm), UV-B (280320 nm), and UV-C (250260 nm)
light. Because (143) has a UV absorption at the UV-BUV-C region, the ease
of photodegradation was UV-C > UV-B > UV-A. Photoinduced dehalogenation
was additionally observed for iodofenphos (141) on glass (Walia et al. 1989b).
Similar photodegradation profiles of chlorpyrifos (145) were more quantitatively
investigated (Walia et al. 1988; Racke 1993). Although the degradation products
were not identified, Meikle et al. (1983) utilized Whatman No. 1 filter paper as
a leaf model and examined rates of either photolysis or volatilization in the
presence of carruba wax. When the aryl moiety is transformed from simple
phenyl rings to heterocycles, similar phototransformation is basically observed.
Oxidation of the P=S moiety concomitant with the P-Oaryl ester cleavage proceeded on glass for quinalophos (148) (Dureja et al. 1988) and pyridafenthion
(150) (Tsao et al. 1989). Although dioxabenzofos (183) possesses a unique cyclic structure, the photoreactions on silica gel obeyed similar mechanisms (Mikami et al. 1977b). Phoxim (152) gave the same photoproducts on glass and
tomato leaf but their chemical identification was not extensively conducted
(Makary et al. 1981).
When 14C- or 32P-phenthoate (168) on glass was exposed to sunlight, about
90% was lost by volatilization after 40 hr, and 20%35% of the residual radioactivity was due to its oxon. The volatilization loss was greatly reduced on silica
gel, with the main product being the oxon (Mikami et al. 1977b). 14C-Sulprofos
(161) was oxidized to the sulfoxide and sulfone similarly as (143) (Ivie and Bull
1976). The basic photochemistry was almost the same for propaphos (159) either on glass or silica gel and was characterized by the significant formation of
the S-oxidized derivatives (Fujii et al. 1979). Direct photolysis is considered
most unlikely for dimethoate (165), malathion (167), phorate (162), and disulfoton (163) not having chromophores in the molecules. The hydrolysis products
with a trace amount of the oxon were detected only for sunlight photolysis of
32
P-(165) (Dauterman et al. 1960). Loss of (167) from the glass surface was
mainly due to thermal vaporization (Awad et al. 1967; El-Refai and Hopkins

Photodegradation of Pesticides

39

1972). Phorate (162) was only photodecomposed to the oxon at 254 nm and
thus sunlight photolysis was considered unlikely (Sharma and Gupta 1994). Insignificant involvement of direct photolysis for (163) was clearly demonstrated
by degradation only at 250260 nm by Hirahara et al. (2001). UV irradiation
of phosalone (170) on glass resulted in dechlorination and cleavage of each
bond at the PSCH2N linkage but without any P=S oxidation (Walia et al.
1989a). Edifenphos (173) exhibited slow degradation under UV irradiation (Ishizuka et al. 1973).
Stereospecific oxidation of the P=S moiety to the oxon was reported for 3HS-2571 (176) on silica gel using UV light (Mikami et al. 1977a). Katagi (1993a)
proposed, in photolysis of butamifos (177), the N-sec-butyl analogue of (176),
the photoinduced intramolecular oxidation of the P=S moiety by the adjacent
nitro group attached to the o-position of the phenyl ring, which may account for
this stereospecificity. Dureja et al. (1989) reported photoinduced oxidation of
the P=S moiety and ester cleavage for isofenphos (175) on glass. 32P-Leptophos
(181) on silica gel was photodegraded with formation of oxon, O-demethylated
and debrominated derivatives, and O-methyl phenyl thiophosphonic acid that
were finally degraded to phenyl phosphonic acid (Zayed et al. 1978). Both oxidation of the P=S moiety and ester cleavage were the main pathways for (181)
on glass (Riskallah et al. 1979) and for cyanofenphos (180) on silica gel (Mikami et al. 1976) under sunlight. These results show the insignificant differences
in photochemistry among phosphorothioates, phosphoramidothioates, and phosphonothioates. Monocrotophos (156) (Dureja 1989; Lindquist and Bull 1967),
phosphamidon (158) (Bull et al. 1967), dichlorvos (153), and trichlorfon (179)
(Derek et al. 1979) do not have any chromophore to absorb sunlight and thus
direct photolysis was most unlikely.
Pyrethroids Chen and Casida (1969) studied photodegradation of the 14Clabels of pyrethrin I (9) and allethrin (10) on glass and identified three possible
oxidation pathways. One of these was the successive oxidation of the transmethyl group of the 2-methylprop-1-enyl moiety via CH2OH and CHO and
finally to the COOH derivative; the other two pathways were formation of transcaronic acid derivative, possibly via ozonide and oxidation at 1-position of the
2-methylprop-1-enyl moiety. Involvement of oxygen species was separately
demonstrated by the increased stability in the presence of an antioxidant such
as dibutylhydroxytoluene (Abe et al. 1972). Isobe et al. (1984) reported the
photoinduced opening of the cyclopropyl ring of (10) followed by either sigmatropic rearrangement to the acrylate or oxidadtive destruction to the mixtures of
glyoxylates. Ruzo et al. (1980) detected the di--methane rearrangement of the
2-(prop-2-enyl)cyclopent-2-enone moiety of S-bioallethrin (12) on glass to form
the cyclopropyl derivatives, but the cis/trans isomerization was minimal. 14CResmethrin (15) on silica gel, glass, or filter paper underwent either epoxidation
of the 2-methylprop-1-enyl moiety or ester cleavage similarly as (10) (Ueda et
al. 1974). These unique reactions were initiated by oxidation of the furan ring,
possibly to give a cyclic ozonide-type peroxide intermediate (Fig. 12). Reduc-

40

T. Katagi

Fig. 12. Photooxidative rearrangement of resmethrin (15).

tion of this intermediate to the diol followed by rearrangement resulted in formation of the cyclopentenolone derivative, especially on silica gel. Migration of
proton or hydrogen radical from the position symmetrical to the benzyl group
or that of the benzyl cation or radical gave the hydroxylactone and benzyloxylactone derivatives, respectively.
Samsonov and Makarov (1996) examined the photostability of (15) on glass
under sunlight and found that the oxidative degradation by O2 was insensitive
to thickness of thin film and likely to be controlled by a filter effect of thin
film. Tetramethrin (23) underwent successive photooxidation at the acid moiety
similar to (10) (Chen and Casida 1969). Ruzo et al. (1982) investigated the
photodegradation of phenothrin (13) and (23) on glass and demonstrated
involvement of oxygen species such as 3O2, 1O2, and O3. The ene reaction of 1O2
(see Fig. 2, reaction 7) was confirmed by formation of 2-methyl-1-hydroperoxyprop-2-ene-1-yl and 2-methyl-1-hydroxyprop-2-ene-1-yl derivatives. For (23),
hydroxylation at the 3-position of the 3,4,5,6-tetrahydrophthaloyl moiety occurred but to a lesser extent. Although the degradation products were not reported, the very rapid photodegradation of cyphenothrin (14) was observed on
silica gel possibly via photoreactions similar to (13) (Dureja et al. 1984). In
addition to the usual photoinduced isomerization in the acid moiety together
with epoxidation, kadethrin (16) underwent opening of the 2-oxo-3-thiacyclopentylidene ring on glass (Ohsawa and Casida 1979). Trace amounts of benzyloxylactone derivatives were detected similarly to (15) but without formation of
the hydroxycyclopentenolone.
Photoinduced cis/trans isomerization mainly proceeded for deltamethrin (22)
on glass under sunlight, and neither photooxidation nor debromination was detected (Ruzo et al. 1977). Maguire (1990) reported the isomerization of 1R-cisisomer to 1S-cis-, 1R-trans-, and 1S-trans forms by chiral HPLC analysis. The

Photodegradation of Pesticides

41

other reactions were ester cleavage to form the corresponding acid, -cyano-3phenoxybenzyl alcohol, and 3-phenoxybenzaldehyde together with decarboxylation (see Fig. 2, reaction 4). The dimethylacrylate derivative was also detected
at a trace amount similar to (10). Cyhalothrin (21) exhibited the similar photoreactions to (22) on glass (Ruzo et al. 1987). Tralomethrin (25) and tralocythrin
(26) were mainly phototransformed to (22) and cypermethrin (19) with their
trans-isomers, respectively (Ruzo and Casida 1981). The 1-bromo derivatives
were also formed possibly via intramolecular abstraction of hydrogen at the 1position by bromine homolytically being photodissociated. Many pyrethroids
have the 3-phenoxybenzyl or -cyano-3-phenoxybenzyl moiety, and the introduction of the cyano group has been demonstrated not to significantly affect
photoreactivity either in solution or on glass (Ruzo and Casida 1982). Holmstead et al. (1978b) reported that the main photo-reaction of fenvalerate (27) on
glass is decarboxylation. Homolytic cleavage of either (C=O)OCH or
C(=O)OCH bond has been evidenced by detection of many relevant photoproducts. By exposure to sunlight, flucythrinate (29) was photodegraded on
glass via almost the same pathways as (27) (Chattopadhyaya and Dureja 1991).
Although the NH linkage was additionally introduced to the acid moiety, almost
the same photoreactions were reported for fluvalinate (30) (Quinstad and Staiger
1984). Sunlight exposure resulted in formation of formanilide, probably due to
reaction of haloanilide with photochemically produced formaldehyde, but no
decarboxylated derivative of (30) was detected on glass. Acrinathrin (31) on
glass was photodegraded to the corresponding acid and alcohol moieties via
ester cleavage (Samsonov and Pokrovskii 2001). The predominant photodegradation process of etofenprox (32) on glass was the photooxidation of the benzyl
moiety to benzoyl, whereas the unique product formed by loss of the CH2O
moiety in solution photochemistry was not detected (Fig. 13). The benzyl radical

Fig. 13. Photodegradation of etofenprox (32).

42

T. Katagi

would be possibly formed by hydrogen abstraction, which reacted with O2 to


the hydroperoxy derivative followed by its dehydration to the benzoyl derivative
(Class et al. 1989; Tsao and Eto 1990b).
Carbamates Substituted phenyl N-methylcarbamates are commonly photodegraded via successive oxidation of the N-methyl group on glass or silica gel
(see Fig. 2, reaction 6) (Abdel-Wahab et al. 1966). Aminocarb (66) or mexacarbate (67) were photodegraded on silica gel at 254 nm to more than 10 photoproducts including the N(CH3)CHO and NHCHO derivatives (Abdel-Wahab and
Casida 1967). Because the same products were detected in photolysis of (66)
on glass at >290 nm, the oxidation process was likely to occur in the environment (Pirisi et al. 2001). This oxidation was also reported for pirimicarb (78)
on glass and cellulose plates, and involvement of OH was assumed (Pirisi et
al. 1996, 1998). Methiocarb (68) underwent photoinduced oxidation of the methylthio group on wax-coated glass to form the corresponding sulfoxide and sulfone, while it was degraded to unknowns on glass (Pirisi et al. 2001). In the
case of benfuracarb (89) on glass, the photocleavage of carbamate linkage at
the NC(=O) bond gave carbofuran (72), which was further photodegraded
to the corresponding phenol (Dureja et al. 1990). At the same time, either
N(CH3)C or N(CH3)S bond was cleaved by irradiation to form N-(2-ethoxycarbonylethyl)-N-isopropylsulfenamine and its derivative (Fig. 14). Photodegradation of benthiocarb (86) gave mainly 4-chlorobenzoic acid via the corresponding alcohol and aldehyde with insignificant S-oxidation (Ishikawa et al. 1977).
Either N-deethylation or hydroxylation at the 2-position of the phenyl ring was
a minor pathway, the latter of which was characteristic of photolysis on glass.
These photoreactions were also observed on silica gel but with a slower photodegradation rate (Cheng and Hwang 1996). Photoinduced homolytic cleavage

Fig. 14. Photodegradation of benfuracarb (89).

Photodegradation of Pesticides

43

of the carbonyl CS bond gave the benzyl thiol and its dimer (Ruzo and Casida
1985). The radicals produced via cleavage of the benzylic CS bond reacted
with O2 to form the series of alcohol-to-acid derivatives or gave benzyl N,Ndiethyl carbamate or benzylamine via release of either sulfur or O=C=S, followed by recombination with the 4-chlorobenzyl radical. N-Methylcarbonyl-Nethyl derivative was also detected as a major product on glass, and involvement
of OH was proposed. Cartap (91) was thought to undergo photocleavage of the
thiocarbamate SC(=O) bond to give the radical, followed by intramolecular
cyclization to nereistoxin on either glass or silica gel (Tsao and Eto 1989). The
photoinduced homolytic cleavage of the carbamate C(=O)O linkage was most
likely for phenmedipham (84) on silica gel by detection of N,N-ditolylurea from
the reactive isocyanate (Emmelin et al. 1998). Schafmeier et al. (1998) reported
that the photolysis of (84) on silica gel was highly dependent on coexisting
humic substances. In the case of fenothiocarb (85), photoinduced oxidation at
sulfur mainly proceeded on silica gel, followed by cleavage of the carbamate
linkage to the very reactive intermediate sulfenic acid that was readily decomposed to 4-phenoxybutylsulfonic acid (Unai and Tomizawa 1986). Thiophanate
methyl (92) was rapidly photodegraded on glass via conversion of the CS
group to ketone and via intramolecular cyclization to MBC (93) (Soeda et al.
1972). MBC (93) was found rather stable on silica gel, but its photodegradation
was accelerated by a sensitizer such as riboflavin (Fleeker and Lacy 1977).
Amides, Anilides, and Dicarboximides Alachlor (34) underwent photoinduced
cleavage of either the NCH2 or NC(=O) bond on glass finally to the corresponding aniline, together with dechlorination and intramolecular cyclization
to the indoline derivative (Fang 1977). Similar photodegradation profiles were
reported for butachlor (36) when irradiated at 254 nm, together with the photosubstitution of Cl with OH followed by intramolecular cyclization to N-2,6diethylphenyl-2,5-dihydrooxazol-4-one (Chen and Chen 1978). These herbicides
possess insignificant absorption at >290 nm and thus the direct photolysis was
considered unlikely in the environment. Either NC(=O) or N-aryl bond of
mepronil (38) was slowly cleaved on silica gel by sunlight to form benzamide
and benzoic acid, followed by stepwise oxidation of the o-methyl group together
with p-hydroxylation at the aniline moiety, but only as a trace amount (Yumita
and Yamamoto 1982). Although photodegradation was found insignificant at 254
nm, flutolanil (39) underwent O-deisopropylation and photo-Fries rearrangement
(see Fig. 2, reaction 9b) to give 2-aminobenzophenone derivative on either silica
gel or glass (Tsao and Eto 1991). The photo-Fries rearrangement of amides at
254 nm has been reported on dry silica gel by Abdal-Malik and de Mayo (1984),
and radical pairs formed were considered not to separate on the surface. The
photolysis mechanism on silica gel was investigated for analogues of (38) having different o-substituents (Yumita et al. 1984). They proposed involvement of
two pathways: one is the NC(=O) bond cleavage with the formed aniline
being polymerized and the other is the hydroxylation at p-position of the aniline
moiety followed by cleavage of the N-aryl bond. Carboxin (42) exhibited rapid

44

T. Katagi

photodegradation on glass, probably via oxidation at the oxathiine sulfur to sulfoxide and sulfone (Buchenauer 1975). Niclosamide (40) underwent extensive
photodegradation either on glass or silica gel to chlorosalicylic acid via cleavage
of the NC(=O) bond (Schultz and Harman 1978). In contrast, naproanilide
(41) exhibited homolytic cleavage of the CHO bond or photooxidation at the
methine carbon to form 2-hydroxypropananilide at >254 nm on glass (Tsao and
Eto 1990a). Unique photoreactions were observed for isoxaben (46) on silica
gel (Mamouni et al. 1992). The N-O bond in the isoxazole ring was first photocleaved and recyclized to the three-membered azirine derivative whose ring was
again opened at the C-C bond to form the oxazole derivative (Fig. 15). It was
considered that these compounds converted to each other under irradiation, and
the azirine and oxazole derivatives were photodegraded to 2,6-dimethoxybenzamide followed by reduction to the corresponding benzonitrile. Concerning
imide pesticides, Sumida et al. (1973) examined the photolysis of experimental
fungicide DDOD (113) on glass but without detailed information on degradates.
Ureas Diflubenzuron (59) was photodegraded on glass or silica gel to 4-chlorophenyl isocyanate and 2,6-dichlorobenzamide, indicating involvement of photoinduced cleavage of the central C-N bond via NH hydrogen abstraction by the
excited-state carbonyl oxygen in the six-membered ring transition state (Ruzo
et al. 1974). Almost half of 14C-(59) was lost from silica gel during 28 d with
trace formation of unknowns (Bull and Ivie 1976). Diafenthiuron (57) was rapidly photodegraded on Teflon sheets to give the corresponding urea through
carbodiimide by reaction with 1O2 (Drabek et al. 1992). Chlorsulfuron (96) on

Fig. 15. Photoinduced rearrangement of isoxaben (46).

Photodegradation of Pesticides

45

silica gel exhibited slow photodegradation to give the corresponding aminotriazine and benzenesulfonamide derivatives via cleavage of the sulfonylurea
bridge (Herrmann et al. 1985). UV spectrum of (96) in methanol did not show
any absorbance at >290 nm and no bathochromic shift was observed by adsorption onto silica gel, implying involvement of indirect photolysis such as the
reaction with OH. Very insignificant photodegradation on glass was reported
for the few sulfonylurea herbicides, and the presence of surfactants in formulation may play a great role in its indirect photolysis in the real environment
(Thomas and Harrison 1990; Harrison and Thomas 1990). Photoinduced cleavage of either bond at the NHC(=O)N(CH3)-triazine moiety or contraction
of the sulfonylurea bridge was reported for tribenuron-methyl (99) on glass
(Bhattacharjee and Dureja 2002).
Azoles and Triazines Rapid photodegradation of triadimefon (188) was observed on glass, and the cleavage either of O-CH or CHC(=O) bond gave the
corresponding 1N-substituted 1,2,4-traizole derivatives (Nag and Dureja 1996).
Especially under UV light, 1-(4-chlorophenoxy)-1-(1H-1,2,4-triazol-1-yl)-2,2dimethylpropane was identified as a main photoproduct, suggesting that a concerted process involving simultaneous combination with loss of CO might proceed rather than simple coupling of the discrete free radicals after loss of the
CO moiety. Triadimenol (189) underwent dechlorination and cleavage of either
CH-Oaryl or C-N bond when irradiated on glass (Clark and Watkins 1986).
Diniconazole-M (190) showed photoinduced E/Z isomerization followed by intramolecular cyclization between 5-positions of the 2,4-dichlorophenyl and
1,2,4-triazolyl rings (Sharma and Chibber 1997). The secondary alcohol moiety
of this cyclized product was oxidized to the corresponding ketone, and successive opening of the hetero ring gave the isoquinoline derivative (Fig. 16). Photoinduced cleavage of the C-triazole bond eliminating 1,2,4-triazole was observed
for propiconazole (191) on glass (Dureja et al. 1987a). Fluotrimazole (195) ex-

Fig. 16. Photoinduced isomerization of diniconazole-M (190) followed by cyclization.

46

T. Katagi

hibited the similar homolytic cleavage of the C-triazole bond on glass followed
by reaction with water to give the corresponding triphenylmethanol (Clark et al.
1983). The photolytic stability of triazine herbicides on glass was briefly studied
in comparison with their solution photolysis (Chen et al. 1984).
Miscellaneous Direct photolysis was found to be of minor importance for 2,4D (1) in accordance with its insignificant UV absorption at >300 nm (Venkatesh
and Harrison 1999). In contrast, the dimethylamino derivative of MCPA (3) was
photochemically produced from the countercation N,N-dimethylamine (Crosby
and Bowers 1985). In the presence of trace water, (3) was considered to be
photodegraded via dechlorination, decarboxylation, and oxidation to many small
molecules. Significant photodegradation of trifluralin (232) by exposure to sunlight was demonstrated by reduced herbicidal activity, but no information on
the degradation pathway was given (Wright and Warren 1965). 14C-Fluchloralin
(236) on silica gel and glass was photodegraded via either release of the chloroethyl group or intramolecular cyclization between the nitro group and 1-position
of the N-propyl moiety to give benzimidazole derivatives together with the unusual formation of the quinoxaline derivative (Nilles and Zabik 1974). This
type of reductive cyclization seems common to dinitroaniline herbicides, as also
demonstrated by photodegradation of pendimethalin (238) on glass (Dureja and
Walia 1989). Rapid photodegradation of isoprothiolane (204) on silica gel involved a dithiolane ring cleavage, ester hydrolysis, decarboxylation, heterocycles formation such as dithietane and trithiolane and sulfur liberation, and finally
the accumulation of S8 was observed (Chou et al. 1980). 14C-Methoprene (246)
quickly dissipated on glass by photolysis and volatilization (Quinstad et al.
1975). The main photoreaction was E/Z isomerization at 2-ene position to give
50 : 50 mixture of (2E,4E) and (2Z,4E) isomers. 7-Methoxycitronellal was identified as one of the main volatiles and considered to be formed via oxidation at
5-position. Sethoxydim (223) was found very unstable, and N-deethoxylated
was the only derivative identified (Campbell and Penner 1985a). Soeda et al.
(1979) investigated the photodegradation of 14C-alloxydim (224) on silica gel
and found that the main pathway was N-deallyloxylation similar to (223), and
(224) also underwent Beckmann rearrangement to give two types of oxo-tetrahydrobenzoxazole derivatives but in lesser amounts.
Bentazone (200) possessing fairly strong absorption at around 300 nm undergoes direct photolysis (Nilles and Zabik 1975). 14C-(200) was lost from a silica
gel surface mainly due to volatilization, whereas photoinduced oxidation of the
NH moiety proceeded on glass followed by elimination of SO2 to give N-isopropyl-o-nitrosobenzoylamine and the successive oxidation to the nitro derivative.
For buprofezin (222) on glass, the few types of ring opening were caused by
sunlight, resulting in formation of ureas and thioureas (Datta and Walia 1997).
Thiabendazole (212) showed very slow sunlight photodegradation on glass via
opening of the thiazole ring to benzimidazole-2-carboxamide and benzimidazole
(Jacob et al. 1975). Photooxidative breakdown of one phenyl ring was observed
for diquat (226) on silica gel, leading to formation of 1,2,3,4-tetrahydro-1-oxo-

Photodegradation of Pesticides

47

pyrido[1,2-a]-5-pyradinium salt and picolinamide (Smith and Grove 1969). The


photodegradation of fipronil (220) on silica gel, glass, and paper gave the desulfinyl derivative with trace amounts of sulfone and sulfide derivatives, possibly
via photoinduced oxidation and reduction (Hainzl and Casida 1996). The desulfinylation was assumed to occur from the hydrogen-bonded six-membered intermediate formed between NH2 and SOCF3 groups (Fig. 17). Perfluidone (206),
having sufficient UV absorption at 330 nm, underwent direct photolysis probably via photochemical breakdown of the CF3SO2NH moiety (Ketchersid and
Merkle 1975). The hydrolytic degradation of 14C-chlordimeform (207) to Nformyl-4-chloro-o-toluidine was accelerated by sunlight exposure on silica gel
(Knowles and Sen Gupta 1969). 14C-Guazatine (245) gave a very odd photoproduct on glass having a carbonyl moiety with a molecular weight greater by
28 than (245) (Sato et al. 1985b). The degradate was considered to be formed
by photooxidation of the methylene group at 4-position to the central NH moiety, followed by methylation at the adjacent methylene carbon. Stepwise dephenylation was observed for fentin acetate (241) on silica gel when exposed to
UV light, finally to an inorganic tin (Barnes et al. 1973). Photooxidation to the
benzoyl derivative was the predominant photodegradation pathway of cinmethylin (249) (Grayson et al. 1987). Imazaquin (230) and imazethapyr (231) photodecomposed on glass and silica gel, probably via decarboxylation or ring decomposition through stepwise oxidation (Basham and Lavy 1987; Goetz et al.
1990; Schroeder 1997).
Azadirachtin-A (253) underwent rapid photoisomerization at the (E)-2-methylbut-2-enoate moiety to the Z-isomer on glass (Dureja and Johnson 2000). At
least 10 primary degradates were formed from avermectin B1a (250) on glass
through sunlight irradiation, but direct photolysis was unlikely due to lack of
UV absorption at >290 nm (Crouch et al. 1991). The main pathway was insertion of oxygen at the 14- or 15-position of the macrocycle with accompanying
double-bond shift, probably via formation of an allylic hydroperoxide from 1O2
by the ene-type reaction. Oxidation at the 8-position, possibly by reaction with
O2 to form the intermediate hydroperoxide, was also observed. For avermectin
B1a analogue MAB1a (251), photolysis on glass was slightly accelerated compared to (250), but no oxidation at the most photosensitive 14- and 15-positions
was observed (Feely et al. 1992). Alternatively, photoinduced isomerization at
8- and 9-positions, oxidative N-dealkylation, and cleavage of the ether linkage
in the sugar region were observed.
D. Photodegradation of Pesticides in Organic Solvents
and Plant Model Systems
Photodegradation profiles of pesticides in organic solvents and thin films of
fatty acids as plant surface models are summarized in Table 8 (see table on page
101). Examples of photodegradation in more complex media such as thin films
prepared from the extracted epicuticular waxes or enzymatically isolated cuticles
are listed in Table 9 (see table on page 104).

Fig. 17. Various photodegradation pathways of fipronil (220).

48
T. Katagi

Photodegradation of Pesticides

49

Organochlorines Photolysis of DDT (130) and methoxychlor (134) was conducted in methyl oleate as a representative of octadecanoic acids in plant cuticles (Schwack 1988). The radicals produced via homolytic dechlorination reacted at the C=C double bond of methyl oleate as well as a chlorine radical to
form stearic acid derivatives and also abstracted hydrogen from methyl oleate
to form the corresponding dichloromethyl derivatives. Dechlorination also occurs at the aryl moiety of anilazine (211) in cyclohexene with irradiation, and
the radicals produced reacted with either solvent molecules or methyl oleate
(Breithaupt and Schwack 2000). Jahn et al. (1999) confirmed the bound formation of chlorothalonil (117) on the enzymatically isolated tomato cuticles under
irradiation by using enzyme-linked immunosorbent assay. Dioxin, 2,3,7,8-TCDD
(129), underwent either dechlorination at 8-position by exposure to UV light or
rearrangement to biphenyl derivative in isooctane (Kieatiwong et al. 1990).
Schuler et al. (1998) reported the photoinduced dechlorination of (129) in a thin
wax film of laurel cherry leaves. Photoinduced homolytic dechlorination followed by reaction with unsaturated C=C bonds was most likely to occur on
plant cuticles. Endosulfan (125) exhibited photoinduced isomerization in hexane
from the -isomer to the -isomer but with loss of two chlorine atoms at the
bridged carbon (Dureja and Mukerjee 1982). Heptachlor (127) gave 1-exohydroxychlordene via substitution of Cl with a hydroxyl group, which underwent either (2 + 2) cycloaddition to the full-cage isomer or intramolecular
ene-reaction, finally forming the cyclic ketone (Parlar et al. 1978).
Organophosphorus Esters Photochemistry of this chemical class has been extensively reviewed by Floer-Muller and Schwack (2001). Schwack (1987) and
Schwack et al. (1994) examined the effect of a reaction medium on photodegradation profiles of parathion (135). UV irradiation of (135) in cyclohexene
caused reduction of the nitro group followed by addition of cyclohexene to the
nitroso group via ene-type reaction. Detection of azo and azoxy dimers of (135)
indicated involvement of stepwise photoreduction of the nitro group to amino
via nitroso and hydroxyamino groups (Fig. 18). The main product in 2-propanol
was the azoxy dimer, part of which rearranged to the 2-hydroxyazo derivative
with no detection of the azo dimer. The amounts of these dimers greatly decreased in cyclohexane containing methyl 12-hydroxystearate (ester of a cutin
acid) with predominant formation of the oxon. These findings imply that the
reaction medium is one of the most important factors controlling the photoprocess of (135), and the proton-donating ability of the solvent molecule may play
a large role. These processes were confirmed by Schynowski and Schwack
(1996) on the enzymatically isolated fruit cuticle of tomato, paprika, apple, and
grape. In the early stage of photolysis, the unstable nitroso derivative of (135)
was dominant, but the 2-hydroxyazo derivative finally gradually accumulated.
The oxon derivative and 4-nitrophenol were minor products. Dissipation rates
of (135) increased with iodine number of the fruit cuticle, showing that the
olefinic portion of cuticles plays an important role in the reaction. In contrast,
oxidation to give sulfoxide and sulfone proceeded for fenthion (143) (Leuch and

Fig. 18. Successive photoreduction of parathion (135) followed by dimerization.

50
T. Katagi

Photodegradation of Pesticides

51

Bowman 1968; Minelli et al. 1996). In the thin film of fruit wax from orange,
nectarine, and olive, (143) gave the sulfoxide under sunlight with a trace amount
of the sulfone (Cabras et al. 1997b; Pirisi et al. 2001). With an increase of wax
thickness, dissipation rates slightly increased for orange and nectarine, whereas
the thickest olive wax gave the slowest degradation. These results suggest that
the component rather than the amount of wax film is one of the important factors
controlling the photodegradation. Fenitrothion (138) mainly underwent photooxidation at the aryl methyl group and the P=S moiety in methanol, hexane, and
acetone (Ohkawa et al. 1974b; Greenhalgh and Marshall 1976). Thiono-thiolo
rearrangement (see Fig. 2, reaction 9a) and denitration were minor pathways.
The difference in reaction profiles between (135) and (138) most likely originates from the adjacent orientation of nitro and methyl groups in (138) that are
known to take an aci-nitro structure via photoexcitation (Katagi 1989). The
homolytic cleavage of a C-I or C-Cl bond together with formation of the oxon
and cleavage of a P-Oaryl or P-S bond was reported for iodofenphos (141)
(Walia et al. 1989b), chlorpyrifos (145) (Walia et al. 1988), and phosalone (170)
(Walia et al. 1989a). The cis/trans isomerization could be detected for tetrachlovinphos (154) in hexane by exposure to UV light (Dureja et al. 1987b). Isofenphos (175) having a different coordination at phosphorus similarly underwent
P=S oxidation to the oxon by UV irradiation in hexane (Dureja et al. 1989).
Pyrethroids Regarding the photolabile chrysanthemic acid moiety, Ruzo and
Casida (1980) examined the excited states involved by using methyl esters of
dichloro- and dibromo-vinyl derivatives. The main reactions in methanol were
cis/trans isomerization via homolytic cleavage of C1C3 bond of the cyclopropane ring to give a biradical followed by recombination together with the reductive dehalogenation in the vinyl moiety. The isomerization is considered to arise
from carbonyl excitation to the excited triplet state, as explained by molecular
orbital calculations (Katagi et al. 1988). By energy transfer experiments using
various sensitizers, the triplet state energy for these chrysanthemates was estimated to be 60 kcal mol1. Insignificant reduction of debromination in the
presence of triplet quenchers (ET = 5053 kcal mol1) showed no involvement
of the triplet state in this process. They also examined the effect of the alcohol
moiety on photochemistry in benzene at 300 nm under sunlight by using several
pyrethroids possessing a 3-phenoxybenzyl, 3-phenylbenzyl, or 3-benzoylbenzyl
group. Intramolecular energy transfer causing enhanced cis/trans isomerization
was likely to be involved (Ruzo and Casida 1982). Photoinduced cis/trans isomerization via triplet excited state was observed for permethrin (17) in organic
solvents, finally giving the cis/trans ratio of 30/70 (Holmstead et al. 1978a).
This type of pyrethroid usually exhibits weak UV absorption around 275 nm,
essentially in n-* character, resulting from the combined transition of the carbonyl system and the lower energy band of the aromatic rings. The ester cleavage
was also the major process forming 3-phenoxybenzyl alcohol and the dichlorovinyl chrysanthemic acid. In addition to the efficient cis/trans isomerization,
cypermethrin (19) in alcohols or aqueous acetonitrile underwent photoinduced

52

T. Katagi

decarboxylation (Ruzo 1983). The cis-isomer in an argon atmosphere simply


gave the corresponding acid and benzyl cyanide while the presence of O2 significantly gave the ketolactone derivative (50%60%; in the case of the transisomer, the caronic acid instead (Fig. 19). Because Rose Bengal in methanol
could not sensitize this oxidative reaction, O3 was likely to participate instead
of 1O2. The profiles of isomerization of deltamethrin (22) was examined in hexane by Maguire (1990), and detailed analysis of the photoproducts in organic
solvents was reported by Ruzo et al. (1977). The higher photoreactivity in methanol than (17) and (19) was considered to stem from more efficient intersystem
crossing by the heavy atom effect of bromine. Photoinduced homolytic cleavage
of the ester linkage was found to occur at either the C(=O)O or OC(CN)
bond. Epimerization at the benzyl carbon occurred in alcohols as a dark reaction
via proton exchange with a solvent molecule. Tralomethrin (25) and tralocythrin
(26) in organic solvents showed similar photodegradation profiles to those on
glass (Ruzo and Casida 1981). Z-cis-Cyhalothrin (21) underwent both E/Z and
cis/trans isomerizations in hexane (Ruzo et al. 1987). The decarboxylated derivative and the corresponding acid and alcohol moieties were the primary products
detected in addition to hexadienes derived from opening of the cyclopropyl ring.
In the case of allethrin (10), di--methane rearrangement via triplet state in the
alcohol moiety (Bullivant and Pattenden 1973, 1976) and epoxidation and oxidation in the acid moiety (Ruzo et al. 1980) proceeded more favorably than
the usual ester cleavage.
Photochemistry of fenvalerate (27) is characterized by a higher efficiency of
decarboxylation (Holmstead et al. 1978b). Mikami et al. (1985a) applied the
spin-trap method in ESR and demonstrated involvement of the two radical species originating from acid and alcohol moieties of (27). For photodegradation
in hexane, no energy transfer from isobutyrophenone and quenching by 1,3cyclohexadiene showed involvement of the long-lived excited singlet state. Almost identical photodegradation profiles have been reported for flucythrinate
(29) in methanol or hexane with maximum degradation at 271 nm, indicating
the involvement of n-* excitation (Chattopadhyaya and Dureja 1991). Etofenprox (32) underwent either photooxidation at the benzyl carbon or photoinduced
homolytic cleavage of the central ether linkage followed by radical recombination to give the derivative with loss of CH2O from (32) (Class et al. 1989; Tsao
and Eto 1990b).
Carbamates The main photodegradation products of xylylcarb (63) and trimethacarb (64) were phenols via cleavage of the OC(=O) bond in ethanol
and cyclohexane (Addison et al. 1974; Kumar et al. 1974). In the case of (63),
the photo-Fries reaction proceeded only in ethanol under irradiation at >265 nm
together with a trace cleavage at the C(aryl)-O bond leading to formation of
xylene. For propoxur (65), photoinduced cleavage of the C(aryl)-O bond occurred in 2-propanol, cyclohexane, and cyclohexene without photo-Fries rearrangement. Contribution from the OC(=O) bond cleavage was less than 1%.
The main product of ethiofencarb (69) in cyclohexane and 2-propanol was the

Fig. 19. Photodegradation of cypermethrin (19).

Photodegradation of Pesticides
53

54

T. Katagi

corresponding sulfoxide with a trace amount of the sulfone (in 2-propanol) (Kopf
and Schwack 1995). A unique reaction, but as a minor pathway, was oxidation
of the benzyl carbon followed by nucleophilic substitution of the thionyl moiety
by the carbamate nitrogen to form 3,4-dihydro-3-methyl-1,3-benzoxazine-2,4dione. Although the simple photoinduced cleavage of the OC(=O) bond was
observed for aminocarb (66) in organic solvents (Addison et al. 1974; Kumar
et al. 1974), oxidative N-demthylation via the N-CHO intermediate proceeded
on glass. However, this oxidation product disappeared in the presence of wax
from nectarine fruits, showing that some wax components may act as scavengers
of radicals (Pirisi et al. 2001). Similar photoreactions were detected for pirimicarb (78) at >280 nm (Schwack and Kopf 1993). In cyclohexane, the intermediate N-CHO derivative of (78) underwent either decarbonylation to the N-demethylated derivative or oxidation of the aryl methyl group at the 6-position.
Stepwise oxidation at the two aryl methyl groups at 5- and 6-positions gave the
unique 5,7-dihydro-5-oxo-furo[3,4-d]pyrimidinyl ring in cyclohexane. Nectarine
(N) and orange (Or) waxes were found to retard the photodegradation of (78)
while the mandarin orange (M) wax accelerated the reaction (Pirisi et al. 1998).
The rate of photolysis could not be correlated with UV absorbance of waxes or
to their amounts, and thus it was considered dependent on the nature of wax.
Because OH would be the promotor of formylation and N-demethylation,
waxes from N and Or should play the role of radical scavengers. In contrast,
the component of M wax was considered to act as a sensitizer. Similar to fenthion (143), the S-methyl sulfur was stepwise oxidized to the corresponding
sulfoxide and sulfone in the thin wax film from nectarine (Pirisi et al. 2001).
These results indicate that wax chemistry is very important in understanding the
photolytic behavior of pesticides after foliar application.
Dicarboximides Direct photolysis is likely to proceed for this class of fungicides having a weak shoulder of UV absorption at >290 nm due to n-* transitions. The most probable reaction is hydrogen abstraction intramolecularly (Norrish type II) or from a solvent molecule by the excited carbonyl oxygen. The
former case is known for N-phthaloylvaline methyl ester (Griesbeck and Gorner
1999) but not for dicarboximide fungicides. The latter reaction has been reported
for photolysis of folpet (107) in cyclohexene (Schwack 1990). The main reaction was the allylic addition of cyclohexene to one of the carbonyl groups to
form the corresponding carbinol with an oxetane formation via the Paterno
Buchi reaction as a minor pathway. As a unique photoreaction, the (4 + 2)1,4-cycloadition at the phenyl moiety to form the benzazepindione derivative
was detected. These reactions with the olefinic carbons indicate photoreactivity
in plant cuticles when (107) is foliarly applied. However, captan (106) exhibited
a different photoreactivity, that is, homolytic cleavage of a C-Cl bond followed
by release of the SCHCl2 group to tetrahydrophthalimide (Schwack and FloerMuller 1990). Procymidone (108), iprodione (109), and vinclozolin (110) commonly underwent photoinduced cleavage of a C-Cl bond. Successive dechlorination of (109) proceeded in 2-propanol (Schwack et al. 1995a). The addition of

Photodegradation of Pesticides

55

a solvent molecule was predominant in cyclohexene whereas mono-dechlorination was the major change in cyclohexane. Procymidone (108) exhibited almost
the same photolytic profiles as (109) (Schwack et al. 1995b). In (110), solvent
addition mainly occurred at the vinyl side chain and successive dechlorination
was observed (Schwack et al. 1995c). This evidence from photodegradation in
organic solvents suggests the possible photoaddition of this class of fungicides
to wax components of plant cuticles.
Azoles Da Silva et al. (2001) revealed through flash photolysis of triadimefon
(188) in cyclohexane, together with analysis of emission spectra, that the first
excited state is n-* localized at the carbonyl group via fast conversion from * state at the 4-chlorophenoxy moiety and detected a 4-chlorophenoxy radical
at 25 nsec after a laser pulse. Product analysis showed involvement of three
main photoreactions (Nag and Dureja 1997). First was cleavage of the PhO-C
bond to give 4-chlorophenol and the corresponding triazole derivative. Either
cleavage of CHC(=O) or C-triazole bond was also observed. The third was
photoreduction of the carbonyl group to form triadimenol (189). Similar profiles
through photolysis in methanol were reported by Clark et al. (1978). Triadimenol (189) having a -* character in the first excited state also underwent
dechlorination and cleavage of CH-C(OH) or C-triazole bond (Clark and Watkins 1986). Although not detected in solution photolysis of (188), the diazirin
derivative via release of the N-CH moiety from 1,2,4-triazolyl ring was formed
from propiconazole (191) in hexane by UV irradiation (Dureja et al. 1987a).
Hexaconazole (192), fluotrimazole (195), and penconazole (193), not having the
carbonyl group are considered to undergo photolysis via -* transition at the
aromatic moieties. Photoliability of (192) was demonstrated in hexane and acetonitrile but without detailed information on degradation (Santoro et al. 2000).
Photoinduced cleavage of the C-triazole bond followed by addition of methanol
molecule was reported for (195) (Clark et al. 1983). Penconazole (193) underwent photocyclization in 2-propanol or cyclohexane between o-position of the
4-chlorophenyl ring and 5-position of the 1,2,4-triazolyl ring to form the 5H,6H(1,2,4-triazole)-[5,1a]-isoquinoline derivative (Schwack and Hartmann 1994).
Similar photocyclization of diniconazole-M (190) was reported by Sharma and
Chibber (1997). Katagi (2002a) examined this photoprocess for the racemic
mixtures of (190) by NMR and molecular orbital calculations and demonstrated
that the reaction proceeds via excited singlet state in a similar manner as reported for cis-stilbene.
Ureas In addition to the usual photoinduced cleavage of an NC(=O) bond
via carbonyl excitation to form the corresponding aniline, oxidative N-demethylation successively proceeded for isoproturon (56) in organic solvents with release of formaldehyde (Kulshrestha and Mukerjee 1986). UV irradiation of
diflubenzuron (59) in methanol caused the cleavage of the central C(=O)
NHC(=O) bond, mainly leading to formation of N-phenyl methylcarbamate and
2,6-difluorobenzamide (Ruzo et al. 1974). Chlorsulfuron (96) and metsulfuron-

56

T. Katagi

methyl (98), which have insignificant UV absorption at >290 nm, are considered
unlikely to undergo direct photolysis in the environment (Yang et al. 1999). In
contrast, chlorimuron-ethyl (102), with UV absorption at 275 nm, was rapidly
photodegraded in methanol and hexane via cleavage of either the N-C ureic or
S-N bond (Choudhury and Dureja 1997b). Contraction of the sulfonylurea bridge
was considered to follow a concerted elimination of SO2 with formation of intermediate radicals being recombined (Fig. 20). In contrast, photolysis in benzene
did not produce this derivative (Choudhury and Dureja 1997c), showing the importance of solvent polarity. Bhattacharjee and Dureja (1999) reported UV photolysis of tribenuron-methyl (99) in organic solvents proceeding via bond cleavage
around the sulfonylurea bridge including its contraction, similar to (102).
Miscellaneous Sunlight photodegradation of pendimethalin (238) was studied
in several organic solvents (Halder et al. 1989). Dureja and Walia (1989) investigated its sunlight photolysis in methanol and found 2,6-dinitro- and 2-amino6-nitro-3,4-xylidene as principal products via N-dealkylation and photoreduction, but no benzimidazole derivatives formed in aqueous photolysis could be
detected. The effect of unsaturated fatty acids and surfactants coexisting in formulation on photolysis of chinomethionat (228) was studied on glass (Nutahara
and Murai 1984). Oleic acid is one of the major fatty acids in leaf extracts of
eggplant and cucumber and accelerated photolysis of (228) as with its selfdecomposition. Similar photoacceleration was observed by addition of the various unsaturated fatty acids or polyoxylene sorbitan oleates (Tween 60, 80, 85).
Because oleic acid by itself was stable under irradiation, some photoreaction
between (228) and the C=C bond of oleic acid similar to the responses of (130),
(134), and (211) might account for these results. Draper and Casida (1985)
reported the ene-reaction of the nitroso derivative photochemically formed in
thin films from nitrofen (216) and CNP (217). The nitroso derivatives reacted
with the C=C bond of various chemicals including methyl oleate to form the
corresponding nitroxides via alkenylarylhydroxyamines being detected by ESR.
The weak ESR signal was observed for beet leaves when treated with (216) in
an extremely high level, showing the ambiguity of this reaction occurring in the
environment. The ESR signal was detected for irradiated methyl oleate film
containing (135) but not for (138), which coincides with dominant formation of
nitroso and its related derivatives from (135) but major detection of the oxidized
derivatives of (138) instead. As a model cuticle instead of organic solvents or
thin film on glass, Caboni et al. (2002) utilized cellulose membrane coated with
epicuticular waxes from olives to evaluate the extent of evaporation, codistillation, and thermodegradation of azadirachtin. In the absence of waxes, codistillation with water proceeded but dissipation of azadirachtin from wax-treated cellulose was not detected.

VIII. Photodegradation of Pesticides on Soil and Clay Surfaces


The photodegradation profiles of pesticides in/on soil and clay, based on the results of the literature survey, are summarized in Table 10 (see table on page 105).

Fig. 20. Photoinduced rearrangement (bridge contraction) of chlorimuron-ethyl (102).

Photodegradation of Pesticides
57

58

T. Katagi

Organochlorines Dichlobenil (116) and chlorothalonil (117) were reported to


undergo photoinduced substitution of Cl with OH or hydration of cyano groups
in aqueous phase although they were rather resistant to photolysis in or on
soils (EPA OPPTS 1998c, 1999c). In contrast, some photodegradation has been
observed for dicamba (118) and chloramben (119). Misra et al. (1997) examined
photodegradation of 14C-(119) by varying conditions of loam soil and found that
3,5-dihydroxy- and 3-chloro-6-hydroxy-benzoic acids are detected in the intact
soil at 75% field moisture while only the deaminated derivative is formed on
air-dried soil. Moisture content might affect the extent of reactive species
formed such as OH, resulting in different product profiles. When (118) coated
on synthetic clay laponite was irradiated with UV light, decarboxylation proceeded in addition to dechlorination and hydroxylation (Aguer et al. 2000). The
photoinduced charge transfer from (118) to clay via interaction of the carbonyl
group with the clay surface was proposed as the reaction mechanism, and residual water molecules on clay were considered to play a role in inducing dechlorination and hydroxylation by polarizing either the O-CH3 or C-Cl bond. Pentachlorophenol (121) rapidly dissipated on soil thin layer by UV irradiation to
form octachlorodibenzo-p-dioxin (Liu et al. 2002). Because detailed examination of photoproducts from (121) as solution or solid has revealed that the photoinduced dechlorination primarily results in formation of free radicals whose successive reactions give dioxins (Piccinini et al. 1998), a similar radical processe
would also operate on irradiated soil surfaces. This mechanism may be accounted for with less formation of dioxins by addition of fulvic acids to soils
that can act as a scavenger of radicals. When 2,3,7,8-TCDD (129) applied to
soils was exposed to sunlight, the degradation, possibly via dechlorination, appeared only in the first 5 days but the coexistence of 1%5% hexadecane maintained constant photodegradation up to 15 d (Kieatiwong et al. 1990). These
results coincided with possible photodegradation near the soil surface and implied that hexadecane might act as a hydrocarbon film allowing solubilization
and migration of (129) from the deeper region of soil to the irradiated surface.
Detection of a trace amount of photodieldrin (124) by soil application of 14Caldrin (122) showed a possible contribution of photodegradation on the soil
surface (Klein et al. 1973). Involvement of photolysis in field dissipation of
organochlorine pesticides including (122) and (123) was demonstrated by monitoring studies in 99 fields (Suzuki and Yamato 1974). DDT (130) in/on soil was
photodegraded under sunlight to DDE (131) via dechlorination (Zayed et al.
1994). Less photoreactivity of dicofol (133) was briefly reported on soil (EPA
OPPTS 1998d).
Organophosphorus Esters Hautala (1978) reported that the apparent quantum
yield of parathion (135) was reduced on soils compared to a solution phase and
that no factors except the amount of soil appeared to correlate with the photolysis rate, indicating the importance of the screening effect by soil. As a main
product of parathion methyl (136), 4-nitrophenol was reported in an outdoor
sunlight photodegradation study (EPA OPPTS 1999e). UV irradiation alone was

Photodegradation of Pesticides

59

ineffective for conversion of (135) to the oxon on soil dusts and clays, but the
existence of O3 at 50300 ppb increased its rate by a factor of 23 (Spencer et
al. 1980). Greater formation of the oxon on soil dusts having less organic matter
clearly demonstrated that the catalytic activity of clay played a large role in
photooxidation. These reaction profiles have been recently confirmed on soil
for 14C-(136) by Kromer et al. (1999). Cyanophos (137) and fenitrothion (138)
were photodegraded on soil thin layers mainly to 4-cyano- or 3-methyl-4-nitrophenol via cleavage of the P-O aryl bond, respectively (Mikami et al. 1976,
1985b). Either the corresponding oxon or O-demethylated derivative was a minor photoproduct from (137) and only the former product was identified for
(138). Fenthion (143) was photodegraded on soil thin layers to 40%80% after
4 days, and insignificant degradation was observed on glass (Gohre and Miller
1986). No information on degradates was available, but product analysis on
disulfoton (163) and methiocarb (68) suggested formation of the corresponding
sulfoxide, most likely via reaction with 1O2 produced on soil surfaces by irradiation. The fastest photodegradation in the soil with the least organic matter content might imply the involvement of clay surface.
Degradation of bromophos (140) and iodofenphos (141) on soil was slightly
enhanced by sunlight irradiation via cleavage of the P-O aryl bonds (Allmaier
and Schmid 1985; Allmaier et al. 1984). Walia et al. (1989b) reported oxidation
of the P=S moiety to oxon and stepwise dehalogenation for (141). 14C-Tolclofos-methyl (142) was photodegraded on soil via either photoinduced oxidation of the P=S moiety to the oxon followed by O-demethylation or cleavage
of the P-O aryl bond to give 2,6-dichloro-4-methylphenol (Mikami et al. 1984b).
More of the oxon was detected for soils with lesser organic matter content,
showing contribution of clay surfaces in oxidation as similarily reported for
(135) and (136). From this aspect, Katagi (1990) investigated the photoinduced
oxidation of (142) in/on clay surfaces. The UV reflectance spectrum of (142)
on kaolinite film exhibited almost the same pattern as that in 10% acetonitrile
with no reflectance above 300 nm. UV light exposure at 320 nm resulted in
rapid degradation of (142) especially for the kaolinite surface where significant
amounts of oxon were formed, clearly demonstrating involvement of indirect
photolysis. Less formation of the oxon was observed when clay was air-dried
or the study was conducted under nitrogen. MS analysis of the oxon formed in
kaolinite treated with H218O showed about 40% incorporation of 18O into the
P=O moiety. Formation of H2O2 was confirmed spectrophotometrically, but the
lack of detection of trans-diacetylacetylene from 2,5-dimethylfuran by HPLC
analysis showed the absence of 1O2 on the irradiated clays. Based on these observations, formation of the oxon could be well explained by reaction with OH
that was produced from the successive reaction with residual water molecules,
with the superoxide anion radical formed probably via photoinduced electron
transfer from clay to O2.
Photodegradation of chlorpyrifos (145) on soil at 254 nm gave the oxon and
3,5,6-trichloropyridinol as main degradates with lesser amounts of dechlorinated
derivatives (Walia et al. 1988). Burkhard and Guth (1979) have shown that

60

T. Katagi

diazinon (144) undergoes photoinduced cleavage of the P-O aryl bond on soil.
Similar results have been reported for sunlight photolysis on sandy loam soil
(EPA OPPTS 2000). Because the aqueous photolysis was of insignificant contribution in its dissipation, some kind of indirect photolysis would operate on
the soil surface. Sunlight photodegradation of quinalphos (148) on soil gave
diquinoxalin-2-thiol, diquinoxalin-2-yl sulfide, and disulfide in addition to quinoxalin-2-ol formed via the usual photoinduced cleavage of the P-O aryl bond
without formation of the oxon (Dureja et al. 1988). The photoinduced thionothiolo rearrangement of (148) followed by cleavage of the P-S aryl bond was
most probable, as reported for photodegradation on clays (Banerjee and Dureja
1999). The photolytic profiles of 14C-dioxabenzofos (183) on soil were found to
be common to other phosphorothioates (Mikami et al. 1977b). Photodegradation
of profenofos (160) on soil gave 4-bromo-2-chlorophenol and O-(4-bromo-2chlorophenyl) O-ethyl O-hydrogen phosphate (Burkhard and Guth 1979). The
latter compound was most likely to be formed by photoinduced hydrolysis, in
contrast to the O-dealkylation of other phosphorothioates.
Qualitatively, the extent of photodegradation of 14C-azinphos-methyl (169)
on soil was found to be reduced with an increase of soil depth and organic
matter (Liang and Lichtenstein 1976). Phosalone (170) mainly underwent cleavage of the SCH2-N bond and was either dechlorinated or oxon derivatives were
detected in trace amounts (Walia et al. 1989a). A similar SCH2-N bond cleavage
was reported for methidathion (171) at >290 nm but at trace amounts (Burkhard
and Guth 1979). On exposure to sunlight, 14C-phenthoate (168) was photodegraded via oxidation, cleavage of P-S or S-C bond, and hydrolysis of carboxylic
ester with formation of the -carboxybenzyl thiol (Mikami et al. 1977b). O,O,STrialkyl derivatives not possessing a chromophore are considered to be degraded
by indirect mechanism if photolysis occurs. Phorate (162) was rapidly degraded
in the field to give its sulfoxide and sulfone (Lichtenstein et al. 1973). Disulfoton (163) showed rapid photodegradation on soil with formation of the corresponding sulfoxide (Gohre and Miller 1986). Based on no effect by sterilization,
reactions of (162) and (163) with 1O2 being formed by sunlight irradiation on
soil were considered most likely. Formation of the corresponding sulfone was
reported for (163) under sunlight (EPA OPPTS 1999d). The photoinduced oxidation of the P=S moiety to the oxon was observed only for ethion (184) (EPA
OPPTS 1995c).
Bensulide (164) exhibited simple photoinduced oxidation of the P=S moiety
by UV irradiation (EPA OPPTS 1999a). Monocrotophos (156) and dicrotophos
(157) are unlikely to undergo photochemical reactions on soil due to lack of
chromophores in their molecules (Lee et al. 1989, 1990). Tetrachlovinphos
(154) was expected to undergo E/Z isomerization, but the main reactions were
O-demthylation and cleavage of the P-O vinyl linkage (Beynon and Wright
1969; Dureja et al. 1987b). Isofenphos (175) underwent photoinduced P=S oxidation to the oxon with cleavage of the P-O aryl bond, but no effect on the PN linkage was observed (Dureja et al. 1989). The same degradation profiles as
aerobic soil metabolism but with a much faster rate were reported on soil (EPA

Photodegradation of Pesticides

61

OPPTS 1998f). Allmaier et al. (1984) reported via soil photolysis of ditalimfos
(178) that the main product was O,O-diethyl phosphoramidothioate, probably
by stepwise hydrolysis of the imide ring, but photoinduced cleavage of the P-N
bond as observed in aqueous photolysis was not detected. The P-C bond of
cyanofenphos (180) remained unaffected by irradiation (Mikami et al. 1976).
Although examples are limited, either the P-N or P-C bond is considered to be
resistant to photolysis on soil.
Phenoxyalkanoates and Esters The apparent quantum yield of methyl ester of
2,4-D (1) on soil was found to be 30 times lower than that in aqueous solution,
which might be accounted for by screening and/or quenching effects of soil
(Hautala 1978). The coarser the soil, the faster the sunlight photodegradation of
mecoprop (4) and its 2,4-dichlorophenyl derivative, which suggests that the
deeper penetration of light into soil facilitated their degradation (Romero et al.
1998). Their declines followed the Hoerl function (y = aebxxc) rather than the
usual first-order kinetics, and increase in moisture content enhanced the photodegradation. Therefore, the transport of these pesticides to the photic zone along
with water movement may control their photodegradation profiles. Norris et al.
(1987) reported that triclopyr (7) was degraded in pastures to give 2-methoxy3,5,6-trichloropyridine and 3,5,6-trichloropyridin-2-ol via successive decarboxylation and O-demethylation.
Pyrethroids Photoinduced isomerization was a minor pathway for soil photolysis of cis- or trans-permethrin (17) (Holmstead et al. 1978a). Sunlight irradiation had no effect on ester cleavage for both isomers, and either dechlorination
or cleavage of the cyclopropyl ring to give dimethylacrylate derivative was also
a minor pathway. Photoinduced isomerization of trans- and cis-cypermethrin
(19) was also found insignificant on soil (Takahashi et al. 1985a). The main
reaction was stepwise hydration of the -cyano group to CONH2 and COOH
and was clearly accelerated by sunlight irradiation, although this hydration was
not affected by irradiation in the other study (EPA FIFRA 1999). Photodegradation of deltamethrin (22) was briefly reported with slight contribution of photolysis on soil, and the main degradation product was the corresponding dibromovinyl chrysanthemic acid (EPA FIFRA 1999). Sunlight photodegradation was
found to be also of less importance for Z-cis-cyhalothrin (21) on soil (Ruzo et
al. 1987). In another 14C study using artificial light, formation of the -CONH2
derivative was reported (EPA FIFRA 1999). In the case of cis-tefluthrin (18),
cis/trans isomerization was reported on soil by UV irradiation although its photodegradation was rather slow (EPA FIFRA 1999). Rapid sunlight photodegradation was reported for 14C-cyfluthrin (20) with formation of 4-fluoro-3phenoxybenzaldehyde via ester hydrolysis and release of cyanide ion from the
corresponding cyanohydrin (EPA FIFRA 1999). Similar to (19), hydration of
the -cyano group to CONH2 predominantly occurred for 14C-fenpropathrin (24)
with more formation under sunlight on soil (Takahashi et al. 1985b).

62

T. Katagi

To clarify the reaction mechanism of photoinduced hydration common to


these pyrethroids, Katagi (1993b) examined effects of moisture content and clay
on photolysis of 14C-(24) using UV light at >290 nm. The degradation rate of
(24) was found to be highly dependent on soil moisture, and UV irradiation
only slightly enhanced the hydration. The amount of -CONH2 derivative significantly increased with decreasing soil moisture, and this reaction was enhanced when clay was used. The shift of C=O vibration in IR spectra to a
higher wavenumber in soil and clay as compared with that in KBr implied some
interaction of the cyano group but not the carbonyl oxygen with these surfaces.
Because clay surface is known to exhibit an extremely high acidity when dried,
the observed hydration of the -cyano group was considered most likely to be
acid catalyzed on clay surfaces. The inconsistency observed for the amount of
-CONH2 derivative of (19) between laboratory and outdoor studies can be
elucidated by the difference of soil moisture; less moisture in soil thin layers
kept outdoors would enhance the hydration.
Fenvalerate (27) dominantly underwent hydration of the -cyano group on
soil with a trace formation of decarboxylated derivative (Mikami et al. 1980).
Katagi (1991) reported hydration of the -cyano group and O-dephenylation of
14
C-esfenvalerate (28) on soil and clays, which was enhanced on exposure to
UV light at >300 nm. It was demonstrated by MS analysis that about half of
18
O was incorporated into the O-dephenylated derivative when photolysis was
conducted on clay prepared from H218O suspension. Because Fentons reagent
was found to give the O-dephenylated derivative of (28), OH photochemically
produced on clay surfaces was most likely to directly react with (28). Furthermore, kaolinite clay was found to catalyze hydration of (28) in the presence of
H2O2, acting as a nucleophile against the cyano carbon to finally give the CONH2 derivative. Flucythrinate (29) and fluvalinate (30) similarly underwent
hydration of the -cyano group and cleavage of the ester (Quinstad and Staiger
1984; Dureja and Chattopadhyay 1995). Most of these pyrethroids have the 3phenoxybenzyl moiety, which is usually released through photodegradation as
PBacid (243). The significant red shift in reflectance spectrum of (243) on kaolinite clay showed possible interaction with the surface. The insignificant spectral overlap the artificial light (>300 nm) implied indirect photolysis; the main
photodegradate on soil and clay was 3-hydroxybenzoic acid (3-HB). The photonucleophilic reaction by OH was unlikely because there was no reaction with
more nucleophilic cyanide ion in aqueous photolysis, and 18O incorporation into
3-HB was demonstrated by MS analysis when aqueous photolysis was carried
out in 50% H218O. Incidentally, Fentons reagent gave 3-HB as a main degradate.
These results indicated that OH photochemically produced on these surfaces
attacked the 3-position of (243) and this ipso substitution resulted in formation
of 3-HB.
Carbamates Contribution of photolysis in dissipation of propoxur (65) on soil
was considered less than that from binding to soil and volatilization (EPA OPPTS 1997b). Methiocarb (68) has a 4-methylthio moiety reactive to 1O2, and the

Photodegradation of Pesticides

63

main photodegradation product was the sulfoxide on soils under sunlight (Gohre
and Miller 1986). UV irradiation of carbaryl (71) at >290 nm on soil showed
less photoreactivity than aqueous photolysis (Hautala 1978). The disappearance
of fluorescence of (71) on soil implied the efficient quenching by soil constituents, which at least in part accounted for reduction of photolysis. The photolysis profiles of carbofuran (72) were briefly examined on field plots as a main
degradate of carbosulfan (88) (Nigg et al. 1984). After an outdoor application,
(88) quickly dissipated on soil to give (72), which was rapidly degraded to give
a trace amount of the 3-keto derivative. Benfuracarb (89) was reported to be
similarly converted to (72) via cleavage of the N(CH3)-S bond on soil by UV
irradiation (Dureja et al. 1990). When 14C-benthiocarb (86) was exposed to sunlight on soil taken from a rice-growing area, the photoinduced S-oxidation to
give the sulfoxide and sulfone was observed similarly as (68) possibly via reaction with 1O2 (Cheng and Hwang 1996). A similar photooxidation of sulfur by
sunlight was reported for butyrate (87) on soil (EPA OPPTS 1993). Molinate
(90) is considered to undergo S-oxidation on soil, but faster degradation under
sunlight was reported without any information on degradation products (Konstantinou et al. 2001). The usual cleavage of carbamate linkage proceeded for
asulam (82) to give sulfanilamide as a main photodegradate on soil (EPA OPPTS 1995a). Desmedipham (83) and phenmedipham (84) have two carbamate
linkages in molecules but the C(=O)O aryl bond was found to be primarily
cleaved. The main pathway of 14C-(83) on soil was formation of ethyl (3-hydroxyphenyl)carbamate, and degradates via photo-Fries rearrangement in aqueous
photolysis could not be detected (EPA OPPTS 1996b). Maneb (94) was quickly
degraded to ETU (95) on soil (Rhodes 1977). After application of 14C-(94) and
14
C-(95) to the ground, the recovered radioactivity from soil rapidly decreased
to form ethyleneurea via S-oxidation of (95). Thiophanate-methyl (92) showed
a similar S-oxidation at one of two C=S moieties on soil exposed to sunlight
together with formation of MBC (93) (EPA OPPTS 2001). Almost quantitaive
formation of (93) was reported for benomyl (81) on soil with 2-aminobenzimidazole as a minor degradate (EPA OPPTS 2001).
Amides, Anilides, and Dicarboximides Soil organic matter was found to accelerate photolysis of propachlor (33) by sunlight irradiation (Konstantinou et al.
2001). For alachlor (34), the rate of photodegradation was enhanced by low soil
organic matter, low pH, or high water content (Fang 1977; Chesters et al. 1989).
Alachlor (34) underwent cleavage of either the NCH2 or NC(=O) bond followed by further degradation to 2,6-diethylaniline and reductive dechlorination
(Fig. 21). As a unique reaction, the intramolecular cyclization occurred between
nitrogen and o-ethyl group to give N-chloroacetyl-7-ethylindoline. Somich et al.
(1988) examined the effect of photolytic ozonation on dissipation of 14C-(34) in
soil and found that O3 caused release of more carbon dioxide from the irradiated
soil. Photolysis of metolachlor (35) on soil by artificial sunlight gave N-chloroacetyl-N-(hydroxyprop-1-en-2-yl)-2-ethyl-6-methylaniline as a main degradate
(see Fig. 21) (Chesters et al. 1989). This product was considered to be formed

Fig. 21. Comparative photodegradation pathways of alachlor (34) and metolachlor (35).

64
T. Katagi

Photodegradation of Pesticides

65

via photoinduced cleavage of the O-CH3 bond. Metalaxyl (37) was slowly photodegraded on soil, but its degradation was found to be also controlled by either
microbial degradation or abiotic factors other than light (Sukul and Spiteller
2001). The slower rate of degradation was observed for soil having a larger clay
content, and light screening by adsorption of (37) into the interlayer of clay
might reduce the effect of irradiation. On air-dried soil, (37) was reported to
undergo cleavage of either the CH2-OCH3 or NC(=O) bond (Saha and Sukul
1997). Under simulated sunlight, carboxin (42) was photodegraded to give the
corresponding sulfoxide, and increased moisture content accelerated photolysis
(Murthy et al. 1998). Because insignificant photodegradation of oxycarboxin
(43) was observed under the same conditions, the main degradation pathway of
(42) was most likely to be photoinduced S-oxidation. Further photodegradation
of the sulfoxide was examined in soil suspension, and oxanilic and malonic
acids were additionally identified (Hustert et al. 1999).
Effects of soil properties such as moisture content and soil depth have been
investigated in conjunction with photolytic profiles of 14C-nicloamide (40)
(Frank et al. 2002; Graebing et al. 2002). The reactive site of (40) was the 4nitro group, which was either reduced to an amino group or substituted with
OH, and these products were finally degraded to 3-chloro-6-hydroxybenzoic
acid. NO3 added as a fertilizer, iron oxide, or humic acid was found to cause
insignificant effects on photodegradation of (40) in moistened soil, but reduced
degradation was observed in air-dried soil. Although the photodegradation pathway was not available, the importance of pesticide transport to soil surface has
been extensively studied for napropamide (47) (Miller and Donaldson 1994;
Donaldson and Miller 1996). The main metabolic pathway of florasulam (48)
in soil was O-demethylation, whose formation was accelerated by sunlight exposure, and this product was further photodegraded via stepwise opening of the
triazolopyrimidine ring (Krieger et al. 2000). Microbial degradation of 14C-(48)
was considered to dominate its dissipation in soil, but one product ASTP (8fluoro-5-methoxy[1,2,4]triazolo-[1,5c]pyrimidine-2-sulfonamide) characteristic
of soil photolysis was formed via cleavage of the NH-aryl bond. Degradation
of captan (106) was slightly enhanced by sunlight exposure on moist soil with
major degradates identified as tetrahydrophthalimide and cyclohex-4-ene-2cyano-1-carboxylic acid (EPA OPPTS 1999b). The former product was considered to be formed via cleavage of the N-S bond, whereas opening of the imide
ring followed by reduction of the amide moiety would also proceed in soil. 14CIprodione (109) exhibited a slightly rapid degradation on soil with formation of
3,5-dichloroaniline and 3-(3,5-dichlorophenyl)-2,4-dioxaimidazolinone (Fig. 22)
(EPA OPPTS 1999b). As a unique degradate, 3-(1-methyl-ethyl)-N-(3,5-dichlorophenyl)-2,4-dioxo-1-imidazolidinecarboxamide was identified, and photoinduced cleavage of the amide linkage in the ring followed by recombination of
the carbonyl radical with the N-isopropyl nitrogen was most likely to be involved. The photochemistry of famoxadone (111) on soil was briefly reported
but the characteristic reaction induced by light exposure was not clarified (Jernberg and Lee 1999).

66

T. Katagi

Fig. 22. Photoinduced rearrangement of iprodione (109).

Ureas Jirkovsky et al. (1997) reported that diuron (53) was photodegraded on
sand via N-demethylation and oxidation to the N-formyl-N-methyl derivative
with a trace formation of monuron (52). As N-demethylation of (53) was demonstrated not to require O2, it was considered that the photoinduced rearrangement of an N-methyl group to carbonyl oxygen gave the corresponding isourea
derivative, which was further hydrolyzed to the NHCH3 derivative. For the Nformyl-N-methyl derivative, the excited carbonyl group was likely to initiate
hydrogen abstraction from the N-methyl group followed by reaction with O2.
Photodegradation of linuron (54) on soil gave NHOCH3, NH2 derivatives, and
3,4-dichloroaniline (EPA OPPTS 1995d). By comparing the results in aerobic
soil metabolism, photolysis was likely to be of minor importance for (54). In
the case of isoproturon (56), the same photoreactions as (53) were observed but
with a slower degradation rate (Kulshrestha and Mukerjee 1986). The photoinduced ring rearrangement was reported for 14C-thidiazuron (58) on soil thin
layer, leading to formation of 1-phenyl-3-(1,2,5-thiadiazol-3-yl)urea (Klehr et
al. 1983). Information on benzoylurea pesticides is very limited, but the results
for photodegradation of diflubenzuron (59) on soil showed an insignificant contribution of photolysis (EPA OPPTS 1997a). Chlorsulfuron (96) was slowly
photodegraded on soil via either O-demethylation or cleavage of the sulfonylurea linkage, but acetylbiuret was not detected (Strek 1998). Similar photodegradation under sunlight was reported for chlorimuron-ethyl (102) together with
unique contraction of the sulfonylurea bridge (see Fig. 20) (Choudhury and
Dureja 1997a). Tribenuron-methyl (99) was photodegraded under sunlight via
N-demethylation with cleavage of either the S-N bond or urea moiety, but any
degradates via bridge contraction were not detected (Bhattacharjee and Dureja
2002). In contrast, photodegradation of 14C-rimsulfuron (103) on soil gave two
major degradates via bridge contraction but no acceleration of degradation by

Photodegradation of Pesticides

67

sunlight exposure was observed (Schneiders et al. 1993). One of the degradates
was N-(pyrimidin-2-yl)-N-(pyridin-2-yl)urea derivative, which was considered
to be degraded to the second by release of the carbamoyl group. Involvement
of direct photolysis on sterilized soil was confirmed for triasulfuron (100) and
thifensulfuron-methyl (105) (Albanis et al. 2002).
Azoles and Triazines Triadimefon (188) underwent photoinduced cleavage of
either O-CH or CHC(=O) bond and reduction of the carbonyl group on soil
under sunlight with the decarbonylated derivative as a minor product (Nag and
Dureja 1996). Judging from its UV absorption, direct photolysis was likely to
proceed and the increase of soil moisture enhanced photodegradation (Murthy
et al. 1998). Propiconazole (191) was photodegraded on soil via cleavage of the
C-triazole bond, liberating 1,2,4-triazole. The unique ring opening was reported
for photodegradation of 14C-prochloraz (240) in both laboratory and field studies
(Hollrigl-Rosta et al. 1999). Prochloraz (240) was gradually photodegraded on
soil to give primarily the formylurea derivative of (240) followed by deformylation (Fig. 23). Atrazine (185) was found resistant to photolysis on air-dried
soils (Curran et al. 1992), although slight sunlight photodegradation of an extent
proportional to the content of soil organic matter was reported (Konstantinou et
al. 2001). Gong et al. (2001) showed that in coarser soil where light can penetrate more deeply, or upon addition of humic acid, or in moistened soil where
movement of the pesticide molecule is facilitated, an increased rate of photodegradation can occur.
Miscellaneous Photolysis seems to be of minor importance for dinitroaniline
herbicides relative to volatilization loss and thermal degradation (Wright and
Warren 1965; Parochetti and Hein 1973; Parochetti and Dec 1978). Although a
detailed photodegradation pathway on soil was not available, photolysis of trifluralin (232) on kaolinite clay was conducted to theoretically investigate the
importance of photic depth in clay and its transport within the layer (Balmer et
al. 2000). On exposure to UV light, (232) on clay was slowly decomposed with
formation of two benzimidazole derivatives. Either photoinduced cyclization between nitro nitrogen and the 1-position of an N-propyl group to give the benzimidazole or N-depropylation was most likely to occur. Cyclization was preceded
by photoinduced reduction of the nitro group to a nitroso derivative whose excited state extracted hydrogen from the C1 position of the N-alkyl group. Photodegradation of fluchloralin (236) on soil gave three benzimidazole derivatives
via similar reactions observed for (232) (Nilles and Zabik 1974). The release
of the chloroethyl moiety was observed with a unique formation of 5-nitro-7trifluoromethyl-1,4-quinoxaline. In photodegradation of ethalfluralin (237), 1Hbenzimidazole-3-oxide was additionally detected (EPA OPPTS 1995b). The existence of this product supported a reaction mechanism where the photoexcited
nitro group extracted hydrogen at the C1 position of the N-alkyl group and radical recombination followed by release of hydroxide resulted in benzimidazoleN-oxide. Insignificant degradation by exposure to sunlight was observed for

Fig. 23. Photoinduced ring opening of prochloraz (240).

68
T. Katagi

Photodegradation of Pesticides

69

butralin (234) on soil (EPA OPPTS 1998b). Halder et al. (1989) reported the
photoinduced N-deethylation of pendimethalin (238) on sunlight-exposed soil;
other reactions such as photoreduction were predominant when exposed to UV
light (Dureja and Walia 1989). In contrast to these dinitroaniline herbicides,
oryzalin (235) was found to be photolabile on soil, forming many unknowns,
and bound 14C via cleavage of the C-N(C3H7)2 bond, hydrolysis of sulfonamide,
and formation of the benzimidazole derivative (EPA OPPTS 1994).
Amitrole (196) was moderately resistant to sunlight photolysis on soil but
underwent C-N bond cleavage to give 1,2,4-triazole (EPA OPPTS 1996a). Another type of photoinduced deamination via cleavage of the N-N bond was reported for metribuzin (201) in an outdoor photolysis study on soil (EPA OPPTS
1998g). Terbacil (197) underwent cleavage of the N-C(CH3)3 bond to give 5chloro-6-methyluracil (EPA OPPTS 1998h). Bentazon (200) underwent oxidative opening of the thiadiazinone ring to give N-isopropyl-2-nitrosobenzamide,
followed by oxidation to the corresponding nitro derivative (Nilles and Zabik
1975). The primary photoprocess was considered to be hydrogen abstraction at
the NH moiety to form the radical whose center would also be located at the 5and 7-position of the benzothiadiazinone ring (o- and p-position toward the
carbonyl moiety) via resonance. Norflurazon (214) showed moderate sunlight
photolysis on soil with N-demethylation being the dominant process (Schroeder
Kvien and Banks 1985). Fipronil (220) exhibited a unique photoreaction to give
the desulfinyl derivative (Bobe et al. 1998b), and the sulfone derivative was
also detected in two field dissipation studies (Bobe et al. 1998a; Fenet et al.
2001). 14C-DTP (221), the herbicidal entity of pyrazolate, mainly underwent
oxidative N-demethylation on paddy soils (Yamaoka et al. 1988).

IX. Photodegradation of Pesticides on Plants


Photodegradation profiles of pesticides on plant surfaces were reviewed based
on surface-wash analysis in plant metabolism studies, unless described otherwise, to distinguish the photochemical conversion from biotic processes. The
plant species and some experimental conditions are summarized in Table 11
(see table on page 120) for each pesticide with its half-life.
Organochlorines Each formulation of DDT (130), aldrin (122), dieldrin (123),
and endrin (126) was sprayed on apple trees in the field and the hexane rinsates
of leaves were analyzed (Harrison et al. 1967). DDT was photodegraded to
DDD (132) and 4,4-dichlorobenzophenone. Aldrin was degraded with concomitant formation of (123). The cage-type ketone and aldehyde derived from reactions at the bridged carbons were detected for (126). When (122), (123), and
(126) were applied to leaves of young bean plant or glass plates under sunlight,
similar photodegradation profiles were identified (Ivie and Casida 1971b; Ivie
et al. 1972). In the same study using apple trees, the -isomer of endosulfan
(125) was found to be a little more persistent than the -isomer with formation
of its sulfate. After application to cotton leaves in the field, the -isomer mainly

70

T. Katagi

underwent dechlorination at the bridge carbon with trace amounts of -isomer


and ring-opening products, whereas the -isomer only gave the latter (Dureja
and Mukerjee 1982). These results agreed with those seen with solution photolysis in hexane. Similar solution photochemistry was also reported for chlordane
(128) applied to cabbage leaves (Parlar et al. 1978). As shown in Fig. 24, N(-trichloromethyl-p-methoxybenzyl)-p-methoxyaniline principally underwent
photoinduced rearrangement on lettuce and bean leaves to N-phenyl--hydroxybenzylamide, which was the common photoproduct in water and on glass or
silica gel (Miller et al. 1974). Direct absorption of light caused a charge transfer
from nitrogen to the CCl3 moiety, followed by the successive release of chloride
ion and recombination to the unstable three-membered intermediate, which was
hydrolyzed to the corresponding amide.
Organophosphorus Esters This class of pesticides undergoes various transformations in plant metabolism, as reviewed by Katagi and Mikami (2000). ElRefai and Hopkins (1966) compared dissipation of parathion (135) topically
applied to garden beans or glass plates. The biphasic decay in contrast to the
simple first-order dissipation on glass implied penetration of (135) into cuticle
and tissue. A similar biphasic foliar decay was observed in residue trials in
orange groves with concomitant formation of the oxon (Popendorf and Leffingwell 1978). Formation and decay of the oxon was found to strongly correlate
with dry and stable weather, and effects of foliar dust and airborne oxidants
were suspected. Spear et al. (1978) conducted residue trials of (135) on dwarf
Eureka lemon trees in the presence or absence of O3 and soil dusts (<50 m).
They showed that the a photoinduced dust-catalyzed process is one of the most
important routes for the foliar P=S oxidation of (135). However, the presence
of soil dust on foliage might retard the decay of the oxon by its stronger adsorption to dust (Adams et al. 1977). Joiner and Baetcke (1973) scrutinized the
degradates from 14C-(135) on cotton. Although the degradates were extracted by
homogenization, the S-ethyl isomer was detected with a trace amount of the Sphenyl isomer, indicating the contribution of photolysis.
Plant metabolism studies using 14C- or 32P-labeled fenitrothion (138) in rice
(Miyamoto and Sato 1965), sugar beet, (Ohkawa et al. 1974b), and apples (Hosokawa and Miyamoto 1974) have been reported. 32P-(138) rapidly penetrated
into tissues of rice plants but information on degradation was not provided.
Analysis of methanol rinses of bean leaves showed that (138) was rapidly degraded via similar pathways as (135), but oxidation of the aryl methyl group to
COOH was characteristically detected. In the case of apple fruits, the oxon and
3-methyl-4-nitrophenol were identified as major degradates in the acetone wash,
the former of which was not detected in the homogenates. In the rinsate, the
corresponding S-isomer was also confirmed. Therefore, both oxon and S-isomer
were most likely to be photochemically produced on the surface of apple fruits,
as was also demonstrated by Fukushima et al. (2003) on tomato fruits. For 14Ccyanophos (137) on bean plants, photolysis was of minor importance (Chiba et
al. 1976). The degradation of bromophos (140) on tomato leaves was also found

Fig. 24. Photodegradation of N-(-trichloromethyl-p-methoxybenzyl)-p-methoxyaniline.

Photodegradation of Pesticides
71

72

T. Katagi

to be less important than metabolism in tissues (Stiasni et al. 1969). In contrast,


fenthion (143) was very rapidly degraded on coastal bermudagrass or corn
plants with successive oxidation of the S-methyl sulfur to form the corresponding sulfoxide and sulfone (Leuch and Bowman 1968). Residue analysis of (143)
and its degradates was conducted for a citrus grove, and both sulfoxide and
sulfone were detected only in the orange peel, showing that these oxidations are
surface reactions on fruits (Minelli et al. 1996). Dechlorination of chlorpyrifos
(145) was confirmed by a trace amount through its photodegradation on the
dorsal leaves of soft shield fern by irradiation, but the main degradates were
3,5,6-trichloropyridinol and the oxon (Walia et al. 1988). Formation of the oxon
from isoxathion (151) on foliage was found insignificant, as demonstrated in
metabolism of 14C-(151) (Ando et al. 1975) and residue trials (Endo et al. 1985).
When phoxim (152) was applied outdoors to tomato plants, the surface rinse of
leaves with acetone was found to contain two unknown photoproducts more
polar than (152) (Makary et al. 1981). The corresponding oxon was confirmed
by GC analysis of corn forage treated with emulsifiable concentrate formulation
of (152) but its route of formation was ambiguous due to the harsh Soxhlet
extraction (Bowman and Leuck 1971). Ivie and Bull (1976) conducted photodegradation studies of 14C-sulprofos (161) on cotton leaves and confirmed the
predominant formation of sulfoxide and sulfone in the methanol rinse. Metabolism of 14C- and 32P-phenthoate (168) in Valencia orange leaves and fruit showed
that the main degradation pathways on the surface were P=S oxidation to the
oxon, stepwise hydrolysis at the S-C bond and carboxylate moiety to mandelic
acid, and hydrolysis of the P-S bond followed by formation of disulfide (Takade
et al. 1976). The other minor reaction, O-demethylation, was detected mostly on
orange tree leaves. Similar degradation profiles were observed for 14C-azinphosmethyl (169) in corn and bean metabolism. Extraction by homogenization only
made a site of conversion difficult to identify, but similar profiles on glass under
irradiation showed possible photodegradation on plant surfaces (Liang and Kichtenstein 1976).
32
P- and 14C-Dimethoate (165) was gradually degraded on leaves via P=S
oxidation followed by conversion of the amide moiety finally to the carboxyl
derivative of the oxon, including oxidative N-dealkylation (Dauterman et al.
1960; Lucier and Menzer 1968). Due to lack of a chromophore in (165), indirect
photolysis may account for a part of these conversions. Formothion (166) underwent rapid conversion to (165) on bean leaves but contribution of photolysis is
unclear (Sauer 1972). Malathion (167) is considered to be also resistant to direct
photolysis, and only hydrolysis of ester linkages is a dominant pathway (Awad
et al. 1967; El-Refai and Hopkins 1972; Mostafa et al. 1974). 14C-Edifenphos
(173) was slowly degraded on rice leaves, presumably via hydrolysis, and thus
the contribution of photolysis also was unlikely (Ishizuka et al. 1973). Photoinduced cis/trans isomerization was expected for mevinphos (155), but its high
volatility made it difficult to confirm this possibility (Casida et al. 1956). Hydrolysis of monocrotophos (156) and dicrotophos (157) was found only predominant
on plant surfaces without any contribution from photolysis (Lindquist and Bull

Photodegradation of Pesticides

73

1967; Beynon and Wright 1972; Bull and Lindquist 1964). Bull et al. (1967)
conducted a metabolism study in cotton using the cis- and trans-isomers of 14Cand 32P-phosphamidon (158) and reported faster degradation of the cis-isomer
on foliage, which would be due to its easier hydrolysis. In contrast, photoinduced E/Z isomerization was observed for 14C-tetrachlorovinphos (154) on
leaves of cabbage, apple, bean, and rice (Beynon and Wright 1969; Dureja et
al. 1987b). Acephate (174), cyanofenphos (180), leptophos (181), and butonate
(182), having P-C and P-N bonds in their molecules, underwent either O-dealkylation or ester hydrolysis, and photolysis on plant surfaces seemed to be of
minor importance (Chiba et al. 1976; Zayed et al. 1978; Bull 1979: Derek et al.
1979).
Pyrethroids The trans- and cis-isomers of 14C-phenothrin (13) were rapidly
degraded on kidney bean and rice plants via unique ozonization of the isobutenyl side chain successively to the corresponding aldehyde and carboxylic acid
(Nambu et al. 1980); this seems to be a typical example of reaction of pesticides
with active oxygen species in air. Measurable isomerization was observed for
the aldehyde and carboxylic acid derivatives, at least in part showing involvement of photoprocess. Instead of oxidation, the cis-isomer of 14C-cypermethrin
(19) predominantly underwent cis/trans isomerization on cotton and bean leaves
grown outdoors (Cole et al. 1982). A similar phototransformation was observed
for deltamethrin (22) (Ruzo and Casida 1979; Maguire 1990). Rapid conversion
of tralomethrin (25) and tralocythrin (26) to (22) and (19) followed by isomerization to the trans-isomers was observed together with a slight epimerization
at the benzyl carbon, possibly as a dark reaction (Cole et al. 1982). Dissipation
of fenpropathrin (24) on leaves of mandarin orange was mostly due to penetration into leaf tissues (Takahashi et al. 1985b). The insignificant contribution by
photolysis was accounted for by the very slight differences in half-lives when
(24) was applied to green beans and tomatoes in winter and spring when sunlight intensity to (24) was different (Martinez Galera et al. 1997). Decarboxylation proceeded for fenvalerate (27) on cotton (Holmstead et al. 1978b) and
bean leaves (Ohkawa et al. 1980). Similar profiles have been separately reported
for metabolism in spring wheat, and the analysis of hexane rinse of leaves
clearly demonstrated that decarboxylation is a photoreaction (Lee et al. 1988).
Similar photoinduced decarboxylation via radical processes was detected for
flucythrinate (29) on French bean leaves with degradates formed via cleavage
at the O-CH bond (Chattopadhyaya and Dureja 1991).
Carbamates Metolcarb (61), xylylcarb (63), and trimethacarb (64) likely dissipated from plant foliage mainly by volatilization, with photochemical reactions
playing a minor role (Slade and Casida 1970; Ohkawa et al. 1974a). Although
the hydroxylated derivative at the methylene carbon of the isobutyl group was
confirmed for 14C-fenobucarb (62) applied to rice leaves, the trace amount detected implied that photolysis was of minor importance (Ogawa et al. 1976).
For propoxur (65), Abdel-Wahab et al. (1966) reported that its foliar application

74

T. Katagi

to garden snapbean resulted in rapid dissipation, but the degradation profiles


were not clear. They also showed stepwise sulfur oxidation at the thiomethyl
group of methiocarb (68) to its sulfoxide and sulfone. On the leaves treated with
aminocarb (66) and mexacarbate (67), the stepwise oxidation proceeded outdoors at the N,N-dimethylamino moiety, evidenced by analysis of chloroform
rinsates (Abdel-Wahab and Casida 1967). Because similar reactions were identified when silica gel was used, the oxidation reactions were likely to be photoinduced processes. Carbaryl (71) underwent hydroxylation at the 4- or 5-position
of the naphthyl moiety and N-methyl group by stem injection (Mumma et al.
1971), but its photolytic behavior on plant surfaces was unclear (Abdel-Wahab
et al. 1966). The contribution of photodegradation was not clear for carbofuran
(72), but either hydrolysis of the carbamate linkage or stepwise oxidation at
the 3-position of the 2,3-dihydrobenzofuranyl moiety proceeded on strawberry
(Archer et al. 1977). When 14C-benfuracarb (89) was applied to leaves of bush
bean, cotton, and corn, (89) was gradually degraded to (72) on leaf surfaces via
cleavage of the N(CH3)-SNRR bond (Tanaka et al. 1985). Cleavage was considered to also occur at the N(CH3)S-NRR bond, resulting in formation of the
dimer derivatives of (72) possessing Sn (n > 1) or derivatives having the chemical structure of (89) but with Sn (n > 1) linkage. Because these products were
detected predominantly on the leaf surface, photodegradation most likely accounted for this conversion. From 14C-carbosulfan (88), similar derivatives having the Sn linkage were detected only in surface rinse of leaves and fruits of
Valencia orange trees (Clay and Fukuto 1984). In addition, S-oxidation to form
the sulfone derivative of (88) occurred mainly on leaf surfaces.
Hydrolysis was the main degradation pathway in plants for oxamyl (74), and
no clear explanation about the extent of photodegradation was given (Harvey et
al. 1978). In contrast, pirimicarb (78) was known to rapidly dissipate from lettuce leaves, not only by volatilization but also by stepwise oxidation of the
N,N-dimethyl group of the pyrimidyl ring, but the extent of the photochemical
contribution was not clear (Cabras et al. 1990). Similar oxidation of the Nmethyl group was reported for fenothiocarb (85) on mandarin orange trees together with S-oxidation to the sulfoxide (Unai et al. 1986). Benomyl (81) and
thiophanate-methyl (92) are known to give the same degradate, MBC (93). MBC
(93) was only formed from (81) on leaves of several plants via release of the
N-butylcarbamoyl group (Baude et al. 1973). 14C-(92) on leaves of apple and
grape trees gave (93) as a main degradate together with the degaradate where
the two C=S groups were oxidized to carbonyls (Soeda et al. 1972). Buchenauer
et al. (1973) examined the photoreactivity of (92) in solid and aqueous phases
using UV light and found that (92) was stable as a solid but rapidly degraded
to (93) in water. Therefore, the photodegradation of (92) on plant foliage seems
to proceed in an aqueous microenvironment, probably originating from morning
dew. MBC (93) was found to be mostly photostable but degradable in the presence of a photosensitizer such as riboflavin (Fleeker and Lacy 1977). Maneb
(94) was rapidly degraded to ETU (95) on leaves of tomato and snapbean plants,
followed by further transformation to either ethyleneurea via C=S oxidation to

Photodegradation of Pesticides

75

C=O or dimerization to Jaffes base (1-(2-imidazoline-2-yl)-2-imidazolinethione) (Rhodes 1977). Similar degradates were identified in aqueous photolysis,
showing that these reactions were likely to proceed via photolysis.
Amides, Imides, and Ureas Flutolanil (39) showed slow dissipation on cucumber leaves via deisopropylation at the phenyl ring followed by methylation or
hydroxylation at the 4-position of the aniline ring (Uchida et al. 1983). Conversion on plant surfaces was minimum for carboxin (42) (Buchenauer 1975). Propyzamide (45) was cyclized via reaction between carbonyl oxygen and ethynyl
carbon to form the oxazoline ring followed by its opening to N-(1,1-dimethylacetonyl)benzamide (Yih and Swithenbank 1971). Because this conversion is
known in its hydrolysis (Katagi 2002b), contribution of photolysis was questionable. Opening of its imide ring was observed for procymidone (108) on cucumber leaves but only in trace amounts (Mikami et al. 1984a). Methazole (112)
underwent either opening of the 1,2,4-oxadiazoline-3,5-dione ring to the 3methyl-1-phenylurea derivative or decarboxylation to the 2-oxo-benzimidazole
derivative, with greater amounts detected on the surface of cotton leaves (Dorough et al. 1973). The latter degradate at least was considered to be formed by
photoreaction because it was also detected in aqueous photolysis. Photoreaction
scarcely proceeded for lenacil (199) on sugar beet (Zhang et al. 1999). Diflubenzuron (59) on cotton leaves was resistant to photolysis with insignificant translocation (Bull and Ivie 1976; Mansager et al. 1979). Rodriguez et al. (2001)
identified 2,6-difluorobenzamide as a sole degradate through residue trials on
pine needles. As cumulative solar irradiation was found to correlate highly with
dissipation of (59), and 4-chloroaniline formed via hydrolysis was not detected,
photodegradation forming the corresponding benzamide and isocyanate was
considered most likely. The lack of detection of the latter degradate may be
accounted for by its high volatility. In contrast, diafenthiuron (57) exhibited
rapid photodegradation on cotton leaves (Drabek et al. 1992) and Chinese cabbage (Keum et al. 2002) via reaction of the thiourea moiety with 1O2 to carbodiimide. Tribenuron methyl (99) underwent cleavage of each bond in the sulfonylurea bridge with its contraction on wheat leaves. These processes were most
likely photoreactions because similar degradates were identified on glass plates
exposed to sunlight (Bhattacharjee and Dureja 2002). In contrast, most of 14Cthifensulfuron (104) remained unchanged in the surface rinse of soybean plants,
with trace amounts of thiophene-2-methoxycarbonyl-3-sulfonamide and the 2amino-1,3,5-triazine derivative (Brown et al. 1993).
Azoles On the leaves of marrow plants, triadimefon (188) was converted to
two diastereomers of triadimenol (189) via reduction of carbonyl group to
CHOH, but the contribution of photolysis was unclear (Clark et al. 1978). Degradation of (189) was briefly investigated on apple leaves with 1-(4-chlorophenoxy)-3,3-dimethylbutan-2-one being detected as the sole metabolite (Clark and
Watkins 1986). Because the same degradate was identified through photolysis
in methanol, its formation on foliage was likely to be a photoprocess. Although

76

T. Katagi

dissipation profiles of the other azole fungicides on grape have been investigated, their degradation pathways are not available (Cabras et al. 1997a, 1998;
Peacock et al. 1994). For fluotrimazole (195), the substitution of the 1,2,4-traizole moiety with the hydroxyl group was confirmed on leaves of barley, with
the photo-induced cleavage of C-N bond being proposed (Clark et al. 1983).
Miscellaneous Although 2,4-D (1) does not possess UV absorption >290 nm,
more of (1) was lost from Zea mays leaves by exposure to UV light >290 nm
than the dark control, and thus the possible photosensitization of either some
components in epicuticular waxes or a coexisting oxysorbic surfactant might
occur (Venkatesh and Harrison 1999). The degradation pathway was not clear
but may be estimated by the results of residue trials for triclopyr (7). When (7)
was sprayed on grasses, it was rapidly degraded to give 3,5,6-trichloro-2-pyridinol with a trace amount of 2-methoxy-3,5,6-trichloropyridine (Norris et al.
1987); this suggests the possible cleavage of O-CH2 and the following decarboxylation are involved in degradation. The photostability of 2,3,7,8-TCDD (129)
was examined on excised leaves of rubber plant under sunlight (Crosby and
Wong 1976). TCDD (129) was degraded possibly via direct photolysis because
its UV absorption maximum was at 300 nm. Although its degradation pathway
was not available, dechlorination was most probable based on the latter photodegradation studies (Schuler et al. 1998). Both dinoseb (202) and dinobuton
(203) were rapidly degraded to several unknown compounds on garden snapbean seedlings, and 5%6% of (203) was found to be converted to (202) (Matsuo and Casida 1970). In the presence of a photosensitizer such as rotenone,
these pesticides were more rapidly decomposed via oxidation, ester cleavage,
and reduction of the nitro group (Bandal and Casida 1972).
Sethoxydim (223) was degraded on grasses to eight unknown products and
the O-deethylated derivative, which were also detected in photolytic and thermal
transformation (Campbell and Penner 1985b). For alloxidim (224), the contribution of photolysis to its dissipation on sugar beet leaves was examined by Soeda
et al. (1979). Alloxidim was degraded via cleavage of the N-O bond to form
the amine derivative and Beckmann rearrangement to the two cyclized derivatives. The former main degradate was also formed either by UV photolysis on
silica gel or by catalytic reduction of (224) with 10% Pd/C in a hydrogen atmosphere, indicating photoinduced reductive dissociation was most probable on the
leaf surface. 14C-Imidacloprid (215) on tomato leaves dissipated under sunlight
via oxidation of the imidazolidinimine ring and stepwise loss of the nitroimino
group to finally form the imidazolidin-2-one derivative (Scholz and Reinhard
1999). As the dark control resulted in minimum degradation, these degradates
originated from photolysis. For fipronil (220), photochemical conversion of the
trifluoromethylsulfinyl moiety via homolytic cleavage of the S-CF3 or S-C(pyrazolyl) bond was confirmed on leaves of corn, sweet pea, and pear (Hainzl and
Casida 1996). In contrast, the main degradate in field residue trials was the
sulfone derivative with a trace of the desulfinyl derivative (Fenet et al. 2001).
The well-aerated leaf surface with higher levels of water under tropical condi-

Photodegradation of Pesticides

77

tions might result in favorable photoinduced S-oxidation. Thiabendazole (212)


was photodecomposed on sugar beet, possibly via oxidation to benzimidazole2-carboxamide and benzimidazole, but only in trace amounts (Jacob et al. 1975).
Photoinduced oxidation of an alkyl chain was reported for guazatine (245) on
dwarf apple trees (Sato et al. 1985a). Oxidation to ketone at the 4-position to
the NH moiety followed by methylation at the 3- or 5-position was proposed
to proceed via reaction with the hydroxyl radical or some sensitizer in wax
components.
Moye et al. (1990) investigated the degradation of 3H-, 14C-avermectin B1a
(250) in celery seedlings. By HPLC analysis of homogenization acetone extracts, the 8,9 isomer formed via photoinduced geometric isomerization at 8and 9-positions of the macrocycle was detected as a main degradate. Analysis
of methanol rinsates from cabbage plant treated with 14C-4-(epi-methylamino)4-deoxyavermectin B1a benzoate [MAB; (251)] showed that surface degradation
was mainly due to photolysis; isomerization of the 8,9-double bond, stepwise
oxidation of the N-methyl group via N-formyl to the amino group, hydroxylation
at the 8-position, and loss of the outer sugar by cleavage of the ether bond
(Wrzesinski et al. 1996). The other example is Spinosad consisting of spinosyn
A and D. Saunders and Bret (1997) utilized 14C-spinosyn A (252) for foliar or
topical application to cotton, turnip, cabbage, and apple fruits. Although no
information on the degradation pathway is available, by analogy with (251),
oxidation at double bonds, N-demthylation, and ether cleavage may occur.

Summary
Photodegradation is an abiotic process in the dissipation of pesticides where
molecular excitation by absorption of light energy results in various organic
reactions, or reactive oxygen species such as OH, O3, and 1O2 specifically or
nonspecifically oxidize the functional groups in a pesticide molecule. In the case
of soil photolysis, the heterogeneity of soil together with soil properties varying
with meteorological conditions makes photolytic processes difficult to understand. In contrast to solution photolysis, where light is attenuated by solid particles, both absorption and emission profiles of a pesticide are modified through
interaction with soil components such as adsorption to clay minerals or solubilization to humic substances. Diffusion of a pesticide molecule results in heterogeneous concentration in soil, and either steric constraint or photoinduced
generation of reactive species under the limited mobility sometimes modifies
degradation mechanisms. Extensive investigations of meteorological effects on
soil moisture and temperature as well as development of an elaborate testing
chamber controlling these factors seems to provide better conditions for researchers to examine the photodegradation of pesticides on soil under conditions
similar to the real environment. However, the mechanistic analysis of photodegradation has just begun, and there still remain many issues to be clarified. For
example, how photoprocesses affect the electronic states of pesticide molecules
on soil or how the reactive oxygen species are generated on soil via interaction

78

T. Katagi

with clay minerals and humic substances should be investigated in greater detail.
From this standpoint, the application of diffuse reflectance spectroscopy and
usage or development of various probes to trap intermediate species is highly
desired. Furthermore, only limited information is yet available on the reactions
of pesticides on soil with atmospheric chemical species. For photodegradation
on plants, the importance of an emission spectrum of the light source near its
surface was clarified. Most photochemical information comes from photolysis
in organic solvents or on glass surfaces and/or plant metabolism studies. Epicuticular waxes may be approximated by long-chain hydrocarbons as a very viscous liquid or solid, but the existing form of pesticide molecules in waxes is
still obscure. Either coexistence of formulation agents or steric constraint in the
rigid medium would cause a change of molecular excitation, deactivation, and
photodegradation mechanisms, which should be further investigated to understand the dissipation profiles of a pesticide in or on crops in the field. A thinlayer system with a coat of epicuticular waxes extracted from leaves or isolated
cuticles has been utilized as a model, but its application has been very limited.
There appear to be gaps in our knowledge about the surface chemistry and
photochemistry of pesticides in both rigid media and plant metabolism. Photodegradation studies, for example, by using these models to eliminate contribution from metabolic conversion as much as possible, should be extensively conducted in conjunction with wax chemistry, with the controlling factors being
clarified. As with soil surfaces, the effects of atmospheric oxidants should also
be investigated. Based on this knowledge, new methods of kinetic analysis or a
device simulating the fate of pesticides on these surfaces could be more rationally developed. Concerning soil photolysis, detailed mechanistic analysis of
the mobility and fate of pesticides together with volatilization from soil surfaces
has been initiated and its spatial distribution with time has been simulated with
reasonable precision on a laboratory scale. Although mechanistic analyses have
been conducted on penetration of pesticides through cuticular waxes, its combination with photodegradation to simulate the real environment is awaiting further investigation.

Photodegradation of Pesticides

79

Table Listing

Table
Table
Table
Table
Table
Table
Table
Table
Table
Table
Table

1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.

............................................................................................................
8084
............................................................................................................
85
............................................................................................................
8687
............................................................................................................
88
............................................................................................................
8991
............................................................................................................
9293
............................................................................................................ 94100
............................................................................................................ 101103
............................................................................................................
104
............................................................................................................ 105119
............................................................................................................ 120128

2,4,5-T
MCPA
Fluazifopbutyl
Pyrethrum-I
Allethrin
Cypermethrin
Esfenvalerate
Alachlor
Metolachlor
Carboxin
Naptalam

Napropamide

Diuron
Flumeturon
Phenmec
Metolcarb
Trimethacarb
Propoxur

47

53
55
60
61
64
65

2,4-D

Pesticide

2
3
6
9
10
19
28
34
35
42
44

No.

328

427

316
329
283
283
287
324

255
294
261
264
268
276

282
293
296

255
250

349
345

340
335
342

230
267
292
292
296

Ex.

309
450
338
332
321

Em.

EPA
A
EPA
EPA
EPA
E

S
M
AW

AW

A
A

A
A
E
E
EPA

Sola

Fluorescence
b

L
r
r
r
r
25

r
r
r

r
r

r
r
25
25
L

Tc

300
282

310
255
261
264
268
280

520
498
375
379
393
400

311

279
282
273
277

280
290
300

Ex.

515
528

462

431
437
481
425

510
495
480

Em.

EPA
EPA
EPA
E

SSW
EPA

SSW
S

EW

E
E
EPA
A

EPA
E
E

Sol
<0.2s
<0.2s

2.7
3.4
2.9
0.8

s
s
s
s

0.2 ms

Phosphorescence

Table 1. Wavelength of fluorescence and phosphorescence spectra of pesticides.

L
L
L
L

L
L
L
L

L
L
L

Sogliero et al. 1985


Moye and Winefordner 1965
Moye and Winefordner 1965
Mueller et al. 1992
Mueller et al. 1992
Bowman and Beroza 1967a
Bowman and Beroza 1967a
Takahashi et al. 1985a
Katagi 1991
Mueller et al. 1992
Mueller et al. 1992
Aaron et al. 1979
Krause 1983
Vannelli and Schulman 1984
Murillo Pulgarin and Garca Bermejo 2002
Argauer 1980
Krause 1983
Vannelli and Schulman 1984
Sogliero et al. 1985
Mueller et al. 1992
Addison et al. 1977
Addison et al. 1977
Addison et al. 1977
Bowman and Beroza 1967b

Reference

80
T. Katagi

Aminocarb

Mexacarbate

Methiocarb

Carbaryl

Carbofuran

Mobam

Dimetilan
Pyrolan
Isolan

67

68

71

72

73

75
76
77

Pesticide

66

No.

Table 1. (Continued).

375
400
365

320

325
304
306

340
326
332
338
334

373
360

375
358

Em.

304
266
270

266

281
280
278

285
281
280
286
285

272
262

262
248

Ex.

E
E
E

AW

E
M
AW

E
EPA
M
AW
EPA

E
EPA

E
EPA

Sola

Fluorescence
b

25
25
25

25
r
r

25
r
r
r
L

25

25
r

25
r

Tc

285

400

290
243
257
260

295
300
282

495,530
485,510
400

460
422
400
421

253
248
290
267
262
285
270
275
288
281

Ex.

460
459
460
430
423
440
430
435
518
467,503

Em.

SSW
E
E
E

SSW
E
E

E
EPA
E
E
EPA
E
E
E
E
EPA

Sol

s
s
s
s
s
s
s
s
s
s

0.6 s
1.6 s
0.7 s

1.6 s

2.0 s
1.7 s

0.6
1.2
0.6
0.6
0.7
0.5
0.2
<0.2
2.1
2.2

Phosphorescence

r
L
L
L

r
L
L

L
L
L
L
L
L
L
L
L
L

Bowman and Beroza 1967b


Addison et al. 1977
Moye and Winefordner 1965
Bowman and Beroza 1967b
Addison et al. 1977
Moye and Winefordner 1965
Bowman and Beroza 1967b
Moye and Winefordner 1965
Bowman and Beroza 1967b
Addison et al. 1977
Argauer 1980
Krause 1983
Sogliero et al. 1985
Vannelli and Schulman 1984
Moye and Winefordner 1965
Bowman and Beroza 1967b
Argauer 1980
Krause 1983
Moye and Winefordner 1965
Krause 1983
Vannelli and Schulman 1984
Bowman and Beroza 1967b
Bowman and Beroza 1967b
Bowman and Beroza 1967b

Reference

Photodegradation of Pesticides
81

Folpet

Dichlorobenil
Chloramben
Picloram
DDT
DDE
DDD

Dicofol
Methoxychlor

Parathion
Fenchlorphos
Diazinon
Chlorpyrifos

116
119
120
130
131
132

133
134

135
139
144
145

Pirimicarb
Propham
Chlorpropham
Benomyl

Pesticide

107

78
79
80
81

No.

Table 1. (Continued).

245

298

285
325
320

310
242

286

Ex.

313
405
425

380
306

300

Em.

EPA

EPA
A
A

M
AW
E
M

Sola

Fluorescence
b

L
r
r

r
r
25
r

Tc

247
295
286
308
305
306
285
320
270
270
245
265
285
270
275
360
300
275
280

402
386
480
440
489
412
450
420
425
426
415
515
386
380
515
475
395
484

Ex.

400

Em.

EPA
E
E
EPA
E
E
EPA
E
E
E
E
EPA

SSW
EW
SSS
EW
SSS
EPA

Sol

s
s
s
s

0.2 s
0.2 s

40ms
0.2 s
0.2 s

2.6 s

0.3 s

0.7
<0.2
<0.1
5.0

Phosphorescence

L
L
L
L
L
L
L
L
L
L
L
L

r
L
25
L
25
L

Argauer 1980
Krause 1983
Bowman and Beroza 1967b
Argauer 1980
Vannelli and Schulman 1984
Aaron et al. 1979
Aaron et al. 1979
Aaron et al. 1979
Aaron et al. 1979
Sogliero et al. 1985
Mueller et al. 1992
Glass 1975
Moye and Winefordner 1965
Moye and Winefordner 1965
Sogliero et al. 1985
Moye and Winefordner 1965
Moye and Winefordner 1965
Sogliero et al. 1985
Moye and Winefordner 1965
Moye and Winefordner 1965
Moye and Winefordner 1965
Moye and Winefordner 1965
Sogliero et al. 1985

Reference

82
T. Katagi

Pesticide

Coumaphos

Pyrazophos
Azinphosmethyl

Phosalone

Phosmet
Triadimefon
Triadimenol
Bentazone

Thiabendazole
Norflurazon
Oxyfluorfen
Ethoxyquin

Paraquat

Diquat

No.

147

149
169

170

172
188
189
200

212
214
218
219

225

226

Table 1. (Continued).

420
320
450
433
340
398
350
440
446
360
345
345

310
320

420
340

380
380
359

Em.

310
285
340
260
310
294
323
360
358
282
285
310

282
287

252
290

320
320
320

Ex.

C
C
M
A
M
A
A
M
AW
A
A
A

M
AW

M
M

M
AW
EPA

Sola

18ns

0.3ps
0.7ps

Fluorescence

21
21
r
r
r
r
r
r
r
r
r
r

r
r

r
r

r
r
L

Tc

411

440

457
420

515
510

Em.

330

305

280
325

325
335

Ex.

EPA

EPA
E

SSW
E

Sol

0.8 s

0.6 s

<0.2 s

Phosphorescence

L
L

T
Argauer 1980
Krause 1983
Sogliero et al. 1985
Vannelli and Schulman
Moye and Winefordner
Argauer 1980
Argauer 1980
Sogliero et al. 1985
Moye and Winefordner
Argauer 1980
Krause 1983
Moye and Winefordner
Da Silva et al. 2001
Da Silva et al. 2001
Argauer 1980
Mueller et al. 1992
Argauer 1980
Mueller et al. 1992
Scrano et al. 1999
Argauer 1980
Krause 1983
Mueller et al. 1992
Villemure et al. 1986
Mueller et al. 1992

1965

1965

1984
1965

Reference

Photodegradation of Pesticides
83

Quinomethionate

Imazaquin
Fenarimol
NAA
Warfarine

228

230
239
247
248

453
360
324
390

380
395

Em.

259
295
290
310

360
362

Ex.

A
E
M
M

M
AW

Sola

0.2 ns

Fluorescence

r
25
r
r

r
r

Tc

295

320

475

360

Ex.

418

535

Em.

SSW

SSW

Sol

Phosphorescence

Argauer 1980
Krause 1983
Vannelli and Schulman 1984
Mueller et al. 1992
Conceicao et al. 1997
Argauer 1980
Argauer 1980
Vannelli and Schulman 1984

Reference

Em. & Ex., emission and excitation wavelengths in nm.


, Not detected.
aSolvent: A (acetonitrile), AW (acetonitrile-water, 1/1), C (cyclohexane), D (diethyl ether), E (ethanol), EW (ethanol-water, 1/9), H (hexane), M (methanol), S (sodium
dodecyl sulfate micelle), EPA (diethyl ether-pentane-ethyl alcohol, 5/5/2), SSW (solid state, Whatman No. 42 paper), SSS (solid state, S&S904 filter paper).
b, lifetime.
cTemperature: r: room temperature; L: 77 K; values in C.

Pesticide

No.

Table 1. (Continued).

84
T. Katagi

2,4-D
Fenpropathrin
Esfenvalerate
Hexachlorobenzene
Pentachlorophenol
Photo-dieldrin
DDT
Parathion
Tolclofos-methyl
Atrazine
Paraquat
NMH
PBacid
Nitrobenzene
N,N-Dimethylaniline
Ethyl pyruvate
PBB
Thiathrene

Crystal violet

1
24
28
115
121
124
130
135
142
185
225
242
243

Aqueous

N.R.
Acetonitrile
Aq. acetonitrile
Hexane
Hexane
Hexane
Hexane
N.R.
Aq. acetonitrile
Methanol
Aqueous
Aqueous
pH 2 & 7 buffer
Cyclohexane
Cyclohexane
Cyclohexane
Hexane
Cyclohexane

Solvent

595

240
278
277
218, 231 (sh)
218, 231 (sh), 305 (br)
193
235, 265 (br)
266
275, 282
226, 268
255
346
292; 279
257
257, 297
338
253
275

max (nm)

N.R.: not reported; sh: shoulder; br: broad; PBB: N-propyl p-benzoylbenzamide.

Pesticide

No.

Absorption
max (nm)

Silica gel
282
Three Japanese soils
280
Kaolinite
276
Silica gel
241, 255, 288
Silica gel/adsorbed
247, 310
Silica gel/adsorbed
264
Silica gel/adsorbed
240, 270 (br)
Silica gel
291
Kaolinite
275, 283
Silica gel
233, 268
5% Na-hectrite suspension 263
Na-montmorillonite Swy-1 360
Silica gel or kaolinite
290; 305
+ Silica gel/slurry
273
+ Silica gel/slurry
238, 276
+ Silica gel/slurry
323
Silica gel
285, 325 8sh9
Na-laponite
265, 290 (sh),
450600 (br)
Na-bentonite
545547

Medium

Reflection

Table 2. Maximum wavelengths of absorption or reflectance spectra of chemicals including pesticides.

Ghosal and Mukherjee 1972

Parlar 1990
Katagi 1993b
Katagi 1991
Gab et al. 1975b
Gab et al. 1975b
Parlar 1980
Parlar 1980
Parlar 1990
Katagi 1990
Frei and Nomura 1968
Bailey and Karickhoff 1973
Margulies et al. 1988
Katagi 1992
Leermakers et al. 1966
Leermakers et al. 1966
Leermakers et al. 1966
Pere et al. 2001
Mao and Thomas 1993

Reference

Photodegradation of Pesticides
85

86

T. Katagi

Table 3. Emission profiles of photosensitizers.


Sensitizer

max

ES

ET

ISC

Reference

Acetone

300

85

78.9

0.90 0.98

Carmichael and Hug


1989; Tsao and
Eto 1994

Acetophenone

330

79

73.9

1.00

Carmichael and Hug


1989; Tsao and
Eto 1994

Xanthone

610

77.6

73.9

1.00

Carmichael and Hug


1989; Ivie and
Casida 1971a
Tsao and Eto 1994

Rotenone

290, 340

63

65

Mallet and Surette


1974; Rau and
Hormann 1981;
Ke et al. 1940

Tryptophan

460

88

64.4 65.8

Carmichael and Hug


1989; SeguraCarretero et al.
2000; NagChaudhuri and
Augenstein 1964

Anthraquinone

390

62.4

0.90

Carmichael and Hug


1989; Ivie and
Casida 1971a

Humic substances

60 62

Zepp 1985; Van


Noort et al. 1988

53 54

TOP-9EO

Tanaka et al. 1991

Riboflavin

440, 470

57.8

50

Tsao and Eto 1994;


Nag-Chaudhuri
and Augenstein
1964; Chambers
and Kearns 1969

Eosin

580

45.4

0.43

Carmichael and Hug


1989

Eosin Y

520

52.5

45.5

Chambers and
Kearns 1969

Rose bengal

44.6

0.8

Tsao and Eto 1994

Rhodamine 6G

620

43.0

0.002

Carmichael and Hug


1989

Rhodamine B

560

49.3

43.0

Chambers and
Kearns 1969

Methylene blue

420

33.0

0.52

Carmichael and Hug


1989

Photodegradation of Pesticides

87

Table 3. (Continued).
Sensitizer

max

ES

ET

ISC

Reference

Chlorophyll b

316, 450

31.1

0.81

Carmichael and Hug


1989

Chlorophyll a

460

29.4

0.53

Carmichael and Hug


1989

max: absorption maximum in nm; ES & ET: energies of excited singlet and triplet states in kcal
mole1; ISC: quantum yield of intersystem crossing; : not applicable; TOP-9EO: nonaethoxylated p-(1,1,3,3-tetramethylbutyl)phenol.

88

T. Katagi

Table 4. Reactions of pesticides with active oxygen species.

No.

Pesticide

Singlet oxygen,1O2
12 S-Bioallethrin
13 Phenothrin
17 Permethrin
23 Tetramethrin
49 Chlorthiamid
86 Benthiocarb
145 Chlorpyrifos
138 Fenitrothion
141 Iodofenphos
146 Potasan
Hydroxyl radical, HO
51 Fenuron
86 Benthiocarb
90
122
136
142

Molinate
Aldrin
Parathion-methyl
Tolclofos-methyl

185 Atrazine

243 PBacid
Ozone, O3
13 Phenothrin
23 Tetramethrin
135 Parathion
136 Parathion-methyl

Source of
active oxygen species
Bengal red B in ethanol, O2, at
360 nm
Rose bengal in acetonitrile, O2,
with sunlamp
Rose bengal in methanol, O2, with
40 W GE lamp
Rose bengal in acetonitrile, O2,
with sunlamp
Riboflavin or methylene blue in
methanol with sunlight
Eosin in methanol with F40BL
fluorescent lamp
Rose bengal in methanol, air with
600 W tungsten lamp
Methylene blue in methanol with
200 W tungsten lamp
Rose bengal in methanol, air with
600 W tungsten lamp
Methylene blue in methanol at
>313 nm

Reference
Ruzo et al. 1980
Ruzo et al. 1982
Holmstead et al. 1978a
Ruzo 1983
Ruzo et al. 1982
Rajasekharan Pillai 1977
Draper and Crosby 1981
Walia et al. 1988
Verma et al. 1991
Walia et al. 1989b
Abdou et al. 1988

Aqueous humic acid suspension at


253.7 nm
Aqueous H2O2 at >285 nm or Fentons reagent
Aqueous H2O2 at >285 nm
Aqueous H2O2 at >285 nm
O3 (85200 ppb) at >290 nm
Clay aqueous suspension at >320
nm
Aqueous H2O2 at >290 nm
Aqueous Fe(ClO4)3 with sunlight
Aqueous semiconductor suspension at >340 nm
Fentons reagent, N2

Aguer and Richard 1996b

Electric discharge, hexane


Electric discharge, hexane
Laboratory ozonizer (30300 ppb)
Micro-ozonizer
Laboratory ozonizer (85200 ppb)

Ruzo et al. 1982


Ruzo et al. 1982
Spencer et al. 1980
Gunther et al. 1970
Kromer et al. 1999

Draper and Crosby


Draper and Crosby
Draper and Crosby
Draper and Crosby
Kromer et al. 1999
Katagi 1990

1981
1984
1984
1984

Sanlaville et al. 1996


Larson et al. 1991
Pelizzetti et al. 1993
Katagi 1992

Eggplant
N.R.
53

Tomato
16 21/CHCl3 3050
[C29,31]

40
[C24,26,28]

Grape
N.R./light
petroleum
1
[C25,27,29,31]

15.0
[C24,26,28]

Lemon
20 40/CHCl3 22.9
[C29,31]

10.2
[C28]

3.2

ALC

12.0
[C24,26,28]

18.9
[C29]

18.2

HC

Orange
30 50/CHCl3 40.1
[C27,29,31]

Apple
989/CHCl3

<Fruits>
Marsh grapefruit
-/CHCl3

Plant species
Quantitya/Extractb

12
[C24,26,28]

43.4
[C24,26,28,30]

28.5
[C24,26,28]

34.7

ALD

47

<0.5

7
[C20,24,26]

18.7
[C24,28,30,32]

19.3
[C26,28,30,32]

69.8
[ursolic]

AC

18
[C46]

ES

Composition (% of wax) (major homologue)

Table 5. Composition of soluble cuticular waxes of fruits and leaves.

30 50 (amyrin)
5 30 (naringenin)

0.8 (ketones)

43.9 (terpenoid)

Others

Baker et al. 1975

Baker et al. 1982

Radler and Horn 1965

Baker et al. 1975

Baker et al. 1975

Belding et al. 1998

McDonald et al. 1993

Reference

Photodegradation of Pesticides
89

6 27
[C29,31,33]

HC

Barley
16.1/CHCl3
1.0
[C29,31,33]

Peach
40 69/CHCl3 10
[C25,27,29,31]
Tea crabapple
30
38/CHCl3
ether (1/1)
[C29,31]
Cherry
5
24/CHCl3
ether (1/1)
[C29,31]
Grape
10
12/CHCl3
ether (1/1)
[C29,31]
Rice
60
0.426/CHCl3
[C33]
Oat
9 21/CHCl3 0.3 6.6

<Leaves>
Bitter orange
9.8/CHCl3

Plant species
Quantitya/Extractb

Table 5. (Continued).

10
[C26,28]
80
[C26,28]

39 65
[C26]

ALD

7.0
[C26]

20
[C26,28,30]

82
[C26]

20
[C26,28,30,32]

31 55
[C26,32,34]

ALC

1.7
[C16,18,26]

2.9 4.9

<5
[C16,18]

70
[ursolic]

30
[ursolic]

20
[ursolic]

18 31
[C16,18]

AC

6.1
[C42,44,46,48]

5
[C42,44,46,48,50]

<5
[C42,44,46]

5
[C40,42,44]

20 30
[C48,50,52]

13 35
[C40,42]

ES

Composition (% of wax) (major homologue)

7
(-diketones)

Trace
(C29,31 ketones)

10 (-amyrin)

10 20 (sterol)

Others

Larsson and Svenningsson 1986


Svenningsson 1988

Bengston et al. 1978

OToole et al. 1979

Baker and Hunt 1981

Baker and Hunt 1981

Baker and Hunt 1981

Bukovac et al. 1979


Baker et al. 1979

Haas and Schonherr 1979

Reference

90
T. Katagi

Spinach
5 10/N.R.

5
[C29,31]

50
[C30]

24.5
[C1834]

>60

70
[C22,24,26]

80
[C26]

14
[C26,28,30,32]

59.4
[C32]

ALC

5
[C26,28,30]

8.6
[C2228]

9
[C28,30,32]

19.5
[C32]

ALD

<5

2.2
[C1834]

<5
[C16]

14
[C26,28,30]

4.8
[C24]

AC

20
[C42,44,46,48]

19.6
[C42,44]

10
[C38,40,42,44]

10
[C44,50]

42
[C44,46]

12.8
[C56]

ES

Composition (% of wax) (major homologue)

5.9 (C23,25 ketones)

5
(-sitosterols)

Trace
(triterpenoids)

4 (sterols)

Others

Baker and Hunt 1981

Sen 1987

Baker 1982

Baker and Hunt 1981

Baker and Hunt 1981

Avato et al. 1990

Barta and Komives 1984

Reference

in g cm2. bOrganic solvent used for extraction. or N.R.: not reported. Composition: HC: hydrocarbon; ALC: alcohol; ALD: aldehyde; AC: acid;
ES: ester.

White clover
16/CHCl3
ether (1/1)

39.0
[C27,29,31]

<5
[C29,31,33]

Sugar beet
3.9/CHCl3
ether (1/1)

Potato
5.4/CHCl3

5
[C29,31,33]

17
[C29,31,33]

3.5
[C31]

HC

Dwarf bean
0.9/CHCl3
ether (1/1)

Maize
N.R./CHCl3

Corn
N.R./CHCl3

Plant species
Quantitya/Extractb

Table 5. (Continued).

Photodegradation of Pesticides
91

92

T. Katagi

Table 6. Diffusion coefficients.


No.

Chemicals

Soil and clay


1 2,4-D

Mediuma

MCb pH OMb C

Db

Reference

Silt loam
(3/69/28)

32.5

6.4 7.1

23 6.33

Scott and Phillips


1973

33 Propachlor

Silt loam
(15/70/15)

23

7.3 2

27 2.28

Ritter et al. 1973

50 Diphenamid

Silt loam
(3/69/28)

38

6.4 7.1

24 2.59

Scott and Phillips


1972

55 Flumeturon

Silt loam
(3/69/28)

32.5

6.4 7.1

23 2.17

Scott and Phillips


1973

25

6.4 7.1

24 6.22

Scott and Phillips


1972

0.58 30 1.57

80 Chlorpropham Silt loam


(3/69/28)
114 Lindane

Silt loam
(//18)

10

123 Dieldrin

Clay loam
(14/19/67)

53RH 7.8 0.2

20 0.051

135 Parathion

Silt loam
(//20)

0.5

25 0.62.9 Gerstl et al. 1979

144 Diazinon

Silt loam
(15/70/15)

23

7.3 2

27 0.48

163 Disulfoton

Silt loam
(//18)

32.8

7.8 2.7* 20 0.13

Graham-Bryce
1969

165 Dimethoate

Silt loam
(//18)

32.8

7.8 2.7* 20 4.94

Graham-Bryce
1969

185 Atrazine

Silt loam
(15/70/15)
Silt loam
(3/69/28)

23

7.3 2

27 1.37

Ritter et al. 1973

38

6.4 7.1

24 3.70

Scott and Phillips


1972

186 Simazine

Silt loam
(3/69/28)

38

6.4 7.1

24 3.28

Scott and Phillips


1972

187 Prometone

Silt loam
(3/69/28)

38

6.4 7.1

24 7.69

Scott and Phillips


1972

194 Triticonazole

Loam clay
(15/54/29)

8.2 1.0

22 3.0

Beigel et al. 1997

232 Trifluralin

Silt loam
(17/66/17)
Silt loam
(3/69/28)
Kaolinite clay

30.4

6.7 4.8

22 0.20

38

6.4 7.1

24 0.52

Jacques and
Harvey 1979
Scott and Phillips
1972
Balmer et al. 2000

Silt loam
(17/66/17)

30.4

6.7 4.8

22 0.05

Jacques and
Harvey 1979

(941/16
67/1051)

727

25 0.87

Sadeghi et al.
1989

235 Oryzalin
Urea

28 0.003

Ehlers et al. 1969a


Farmer and Jensen
1970

Ritter et al. 1973

Photodegradation of Pesticides

93

Table 6. (Continued).
No.

Chemicals
PEG4000

Mediuma
Sandy loam
(38/47/15)

p-Nitroanisole Kaolinite
Waxes and cuticles
1 2,4-D

Barley waxes
Isolated citrus
cuticles

MCb pH OMb C
3050 2.5

Db

Reference

25 0.31.6 Barraclough and


Nye 1979
28 0.0069

Balmer et al. 2000

25 1 106 Schreiber and


Schonherr 1993
3 104 Schonherr and
Riederer 1989
25 2 106 Schreiber and
Schonherr 1993

121 Pentachloropenol

Barley waxes

185 Atrazine

Isolated citrus
cuticles

189 Triadimenol

Barley waxes

25 4 107 Schreiber and


Schonherr 1993

Water

23 51.8

Water
Eight
pesticides
aMedium:

9 105 Schonherr and


Riederer 1989

Scott and Phillips


1973

values in the parentheses are weight % of sand, silt, and clay in soil.
soil moisture content in %; RH: relative humidity; OM: soil organic matter content in % (*,
soil organic carbon content); D: apparent diffusion coefficient in mm2 day1.
bMC:

94

T. Katagi

Table 7. Photodegradation of pesticides on glass and silica gel surfaces.

No.

Pesticide

Carrier (Label),
App DT50
(light/dark)

Light source
(wavelength or
season, filter)

Reference

1 2,4-D

GL(14C)
N.R.

UV fluorescent lamp
(max. 356 nm)

Venkatesh and
Harrison 1999

3 MCPA

GL, 6 103
2.5 d/N.R.

Sunlight (summer)

Crosby and Bowers 1985

GL (14C), 2.6
<5 hr/>32 hr
GL, 300
N.R.

275W G.E. sunlamp

Chen and Casida


1969
Kimmel et al. 1982

10 Allethrin

GL, 4 103
N.R.
13 Phenothrin

RPR 3500 UV lamp


(> 290 nm, Pyrex
glass)
15W Fluorescence
lamp

Isobe et al. 1984

GL, 100300
N.R.
GL (10EC), 30
1000
3 hr/N.R.

Sunlight (Sept.)

Ruzo et al. 1982

Sunlight (summer)

Samsonov and
Makarov 1996

14 Cyphenothrin

SG (14C)
2 hr/N.R.

RPR 3500A UV lamp


(max. 360 nm)

Dureja et al. 1984

15 Resmethrin

SG (14C), 1017
N.R.

Sunlight

Ueda et al. 1974

16 Kadethrin

GL (14C), 35
N.R.

Sunlight

Ohsawa and
Casida 1979

22 Deltamethrin

GL (14C), 40
N.R.

Sunlight

Ruzo et al. 1977

23 Tetramethrin

GL (14C), 2.6
<5 hr/>32 hr
GL, 100300
N.R.

275W G.E. sunlamp

Chen and Casida


1969
Ruzo et al. 1982

25 Tralomethrin

GL (14C), 100
N.R.

Sunlight

Ruzo and Casida


1981

26 Tralocythrin

GL (14C), 100
N.R.

Sunlight

Ruzo and Casida


1981

27 Fenvalerate

GL, 127
4d

Sunlight (JulyAug.,
Pyrex glass)

Holmstead et al.
1978b

29 Flucythrinate

GL
11.8 hr/N.R.

UV light in a Rayonet
reactor

Chattopadhyaya
and Dureja 1991

30 Fulvalinate

GL (14C), 24
1 d/N.R.

Sunlight, outdoors

Quinstad and
Staiger 1984

Sunlight (Sept.)

Photodegradation of Pesticides

95

Table 7. (Continued).

No.

Pesticide

Carrier (Label),
App DT50
(light/dark)

31 Acrinathrin

GL, 0.310

32 Etofenprox

GL, 14
1.7 d

2.7 hr/N.R.

GL, 140
1.94 hr

Light source
(wavelength or
season, filter)

Reference

500 W high-pressure
Hg lamp
(313 nm, glass filter)

Samsonov and
Pokrovskii 2001

RPR 3000 UV lamp


(290320 nm, Pyrex
glass)
GL10 UV lamp (254
nm)

Class et al. 1989

Tsao and Eto


1990b

34 Alachlor

GL
6 hr/N.R.

Sunlight (Mar.Apr.)

Fang 1977

36 Butachlor

GL
1.5 hr/N.R.

Germicidal lamp (254


nm)

Chen and Chen


1978

38 Mepronil

SG (14C)
36 d/N.R.
SG, 25
N.R.

Sunlight (Sept.Dec.)

Yumita and Yamamoto 1982


Yumita et al. 1984

400W high-pressure
Hg lamp (max. 365
nm)

39 Flutolanil

GL, 2.5 103


N.R.

Germicidal UV lamp
(254 nm)

Tsao and Eto 1991

40 Niclosamide

SG (14C)
20.5 hr/>7 d

Long-wave lamp
(290405 nm)

Schultz and Harman 1978

41 Naproanilide

GL
N.R.

UV germicidal GL10
lamp (>254 nm)

Tsao and Eto


1990a

42 Carboxin

GL
10 hr/N.R.

Sunlight

Buchenauer 1975

46 Isoxaben

SG
N.R.

Xenon lamp (Hereaus


Suntest)

Mamouni et al.
1992

57 Diafenthiuron

TF
1 hr/N.R.

UV light

Drabek et al. 1992

RUL 3000 lamp (max.


300 nm)
Sunlight (summer)

Ruzo et al. 1974

59 Diflubenzuron GL, SG
N.R.
SG(14C), 4.0
4 wk/N.R.

Bull and Ivie 1976

66 Aminocarb

SG (14C), 400
N.R.

UV light at 253.7 nm

Abdel-Wahab and
Casida 1967

67 Mexacarbate

SG (14C, 3H), 400


N.R.

UV light at 253.7 nm

Abdel-Wahab and
Casida 1967

96

T. Katagi

Table 7. (Continued).

No.

Pesticide

78 Pirimicarb

Carrier (Label),
App DT50
(light/dark)
GL
64 min
CE, 0.03
19 min

Light source
(wavelength or
season, filter)

Reference

125W high-pressure
Hg lamp (> 290
nm, Pyrex glass)
Sunlight (Feb.), outdoors

Pirisi et al. 1996

84 Phenmedipham

SG
15 d/144 d

UV light from Xe
lamp (Heraeus suntest)

Schafmeier et al.
1998

85 Fenothiocarb

SG (14C)
45 hr/N.R.

Sunlight (Sept.Oct.)

Unai and Tomizawa 1986

86 Benthiocarb

GL (14C)
1.7 hr/N.R.

Ishikawa et al.
1977

GL (14C), 1 104
N.R.

High-pressure Hg
lamp (max. 365
nm)
Sunlight (3.54.5 mW
cm2)

89 Benfuracarb

GL
N.R.

Low-pressure Hg
lamp (254 nm)

Dureja et al. 1990

91 Cartap

GL, 12.7
N.R.

Germicidal lamp (373


nm)

Tsao and Eto 1989

92 Thiophanatemethyl

GL (14C), 33
2.8 d/N.R.

Sunlight (MayAug.),
outdoors

Soeda et al. 1972

93 MBC

SG (3T,14C), 5.5
N.R.

Sunlight (July)

Fleeker and Lacy


1977

96 Chlorsulfuron

SG
60 hr/N.R.

125W high-pressure
Hg lamp (>290 nm,
borosilicate glass)

Herrmann et al.
1985

99 Tribenuronmethyl

GL
7 d or 11 hr/N.R.

Sunlight (May) or UV
light (max. 254 nm)

Bhattacharjee and
Dureja 2002

GL (14C), 6.6

500 W Xenon arc


lamp

Sumida et al. 1973

113 DDOD

Cheng and Hwang


1996

8 d/>20 d
121 Pentachlorophenol

GL, 1 103
N.R.

125 W high-pressure
Hg lamp (>290 nm,
water filter)

Piccinini et al.
1998

122 Aldrin

GL
N.R.
SG
N.R.

Sunlight (JuneJuly)

Rosen and Sutherland 1967


Gab et al. 1975a

125 W high-pressure
Hg lamp (>290 nm,
Pyrex glass)

Photodegradation of Pesticides

97

Table 7. (Continued).

No.

Pesticide

Carrier (Label),
App DT50
(light/dark)

Light source
(wavelength or
season, filter)

Reference

123 Dieldrin

GL
1 hr/N.R.

GE G30T8 germicidal
lamp

Benson 1971

125 Endosulfan

GL
N.R.

Sunlight (Mar.)

Dureja and Mukerjee 1982

128 Chlordane

GL, 890
N.R.
SG
N.R.

Sunlight (summer)

Benson et al. 1971

125 W high-pressure
Hg lamp (>290 nm,
Pyrex glass)

Gab et al. 1975a

130 DDT

GL (14C)
N.R.

15W germicidal lamp


(254 nm)

Mosier et al. 1969

136 Parathionmethyl

GL (14C), 0.84
1 d/>1 d

Xe lamp in Suntest
CPS+ (>290 nm,
volatility chamber)

Kromer et al. 1999

137 Cyanophos

SG (14C)
4 d/N.R.

Sunlight (Sept.Oct.),
outdoors

Mikami et al. 1976

138 Fenitrothion

SG (14C)
15 min6 d/N.R.

172.5 W high-pressure
Hg lamp or sunlight
(Nov.)

Ohkawa et al.
1974b

141 Iodofenphos

GL
N.R.

1 kW metal halide
lamp (Applied Photophysics 9500)

Walia et al. 1989b

143 Fenthion

GL
0.021.34 hr/N.R.

Fluorescent lamp
(380750 nm), UV
light (UV-A, UV-B,
UV-C)

Hirahara et al.
2001

145 Chlorpyrifos

GL
18.7 d/N.R.
FP (14C), 1.18
3.2 d/N.R.
GL
N.R.

Low-pressure Hg
lamp (254 nm)
UV light (Pyrex glass)

Walia et al. 1988

148 Quinalphos

Meikle et al. 1983

Low-pressure UV
lamp (254 nm)

Dureja et al. 1988

150 Pyridafenthion GL, 6.4


N.R.

Black light fluorescence lamp (>300


nm)

Tsao et al. 1989

152 Phoxim

UV light (254 nm &


350 nm, uncovered
Petri dish)

Makary et al. 1981

GL
419 hr/N.R.

98

T. Katagi

Table 7. (Continued).

No.

Pesticide

Carrier (Label),
App DT50
(light/dark)

Light source
(wavelength or
season, filter)

Reference

156 Monocrotophos

GL (32P)
5.5 d/N.R.
GL, 1.6
N.R.

Sunlight in a greenhouse.
15W germicidal lamp
(253.7 nm) or sunlight (Apr.May)

Lindquist and Bull


1967
Dureja 1989

159 Propaphos

SG, GL (14C)
1.2 hr/>72 hr

Sunlight (Apr.Aug.)

Fujii et al. 1979

161 Sulprofos

GL (14C), 53
1.1 d/N.R.
GL
5 hr/>13 hr

Sunlight (summer)

Ivie and Bull 1976

Germicidal lamp (254


nm)

Sharma and Gupta


1994

162 Phorate
163 Disulfoton

GL
0.4143.3 hr/N.R.

Fluorescent lamp
(380750 nm), UV
light (UV-A, UV-B,
UV-C)

Hirahara et al.
2001

165 Dimethoate

GL (32P)
N.R.

Sunlight in a greenhouse.

Dauterman et al.
1960

167 Malathion

GL (F)
N.R.

UV (253.7 nm) or
fluorescent (366
nm) light

Awad et al. 1967

168 Phenthoate

GL (14C,32P)
<3 d/N.R.
SG (14C)
4 d/>8 d

Sunlight (Mar. &


Sept.)
Sunlight (Sept.Oct.),
outdoors

Takade et al. 1976

170 Phosalone

GL
8.5 d/>15 d

1 kW high-pressure
metal-halide lamp
(>300 nm)

Walia et al. 1989a

173 Edifenphos

GL (32P), 5.0
10 d/N.R.

UV light

Ishizuka et al.
1973

175 Isofenphos

GL
N.R.

Low-pressure Hg
lamp

Dureja et al. 1989

176 S-2571

SG (3T), 300
N.R.

172.5W high-pressure
Hg lamp

Mikami et al.
1977a

Sunlight (Sept.Oct.),
outdoors

Mikami et al.1976

UV light (310 nm)

Zayed et al. 1978

Sunlight

Riskallah et al.
1979

180 Cyanofenphos SG (14C)


2 d/N.R.
181 Leptophos

SG (32P)
3.4 d/N.R.
GL, 100
20 d/N.R.

Mikami et al.
1977b

Photodegradation of Pesticides

99

Table 7. (Continued).

No.

Pesticide

Carrier (Label),
App DT50
(light/dark)

Light source
(wavelength or
season, filter)

Reference

183 Dioxabenzofos

SG (14C)
2 d/>8 d

Sunlight (Sept.Oct.),
outdoors

Mikami et al.
1977b

188 Triadimefon

GL, 32
1.32.8 hr/N.R.

UV light (254 nm) or


sunlight (Sept.)

Nag and Dureja


1996

189 Triadimenol

GL, 51
N.R.

400 W medium-presClark and Watkins


sure Hg lamp (boro1986
silicate glass)

190 Diniconazole- GL, 50


M
N.R.

Sunlight (June)

Sharma and Chibber 1997

191 Propiconazole GL, 147


N.R.
195 Fluotrimazole GL, 13-50
N.R.

Sunlight

Dureja et al. 1987a

100 W medium-presClark et al. 1983


sure Hg lamp (borosilicate) or summer
sunlight

GL (14C)
>5 d/N.R.

UV lamps (300 & 350


nm)

Nilles and Zabik


1975

204 Isoprothiolane SG
3 hr/N.R.

10 W germicidal lamp
(254 nm)

Chou et al. 1980

206 Perfluidone

GL, 33.6
36 wk/N.R.

UV light (254 or 365


nm)

Ketchersid and
Merkle 1975

207 Chlordimeform

SG (14C)
N.R.

Sunlight

Knowles and Sen


Gupta 1969

212 Thiabendazole GL (14C), 19


>4 mon/N.R.

Sunlight in a greenhouse

Jacob et al. 1975

220 Fipronil

SG, GL, FP
N.R.

Sunlight

Hainzl and Casida


1996

222 Buprofezin

GL
15 d/N.R.

Sunlight (Feb.Mar.)

Datta and Walia


1997

223 Sethoxydim

GL (14C)
<1 hr/N.R.

Sunlight in a greenhouse (350 E m2


sec1)

Campbell and Penner 1985a

224 Alloxydim

SG (14C)
0.7, 4.4 hr/ N.R.

12W UV light (253.7


or 365 nm)

Soeda et al. 1979

226 Diquat

SG (14C)
N.R.

Sunlight (May June),


outdoors

Smith and Grove


1969

232 Trifluralin

GL
N.R.

Sunlight (JuneJuly)

Wright and Warren


1965

200 Bentazone

100

T. Katagi

Table 7. (Continued).

No.

Pesticide

236 Fluchloralin

Carrier (Label),
App DT50
(light/dark)
GL (14C)
48 hr/N.R.

SG (14C)
N.R.
238 Pendimethalin GL
N.R.

Light source
(wavelength or
season, filter)
RPR UV light (300 &
350 nm, Pyrex
glass)

Reference
Nilles and Zabik
1974

Sunlight (Aug.Oct.)
Low-pressure Hg
lamp (254 nm)

Dureja and Walia


1989

241 Fentin acetate

SG (14C)
N.R.

125W high-pressure
Hg lamp (max. 365
nm)

Barns et al. 1973

245 Guazatine

GL (14C), 0.5
36 hr/N.R.

Sunlight lamp (950


E m2 sec1)

Sato et al. 1985b

246 Methoprene

GL (14C), 11
6 hr/N.R.

Sunlight (Oct.)

Quinstad et al.
1975

249 Cinmethylin

GL, 2 103
N.R.

1 kW Xe lamp (Oriel
solar simulator,
AM1 filter)

Grayson et al.
1987

250 Avermectin
B1a

GL, 0.7
23 hr/N.R.

Sunlight

Crouch et al. 1991

251 MAB1a

GL, 0.7
6.2 hr/N.R.

275 W Suntanner RS
bulb (6070 mW
cm2 h1)

Feely et al. 1992

253 AzadirachtinA

GL
48 min/N.R.

UV light (254 nm)

Dureja and Johnson 2000

Medium: Thin film of a pesticide is basically prepared from its organic solution followed by vaporization of solvent. When unspecified, nonradiolabeled pesticide was used.
F, formulation. Materials of carrier are glass (GL), silica gel (SG), filter paper (FP), cellulose
sheet (CE), and Teflon sheet (TF). Label, radiolabel; App, application rate in g cm2 if described
in the literature; N.R., not reported.

Photodegradation of Pesticides

101

Table 8. Photodegradation of pesticides in organic solvent as model plant cuticles.

No.

Pesticide

Medium
DT50

Light source
(wavelength or
season, UV filter)

Reference

17 Permethrin

Methanol
11.5 hr

RPR 3000 lamp


(290320 nm)

Holmstead et al.
1978a

21 Cyhalothrin

Cyclohexane
N.R.

RPR 3000 UV lamp


(290320 nm,
Pyrex glass)

Ruzo et al. 1987

22 Deltamethrin

Hexane
2 days

Summer sunlight
(Pyrex glass, outdoors)
Sunlight

Maguire 1990

Hexane
N.R.

Ruzo et al. 1977

27 Fenvalerate

Hexane
18 min

RPR 3000 UV lamp


(290320 nm,
Pyrex glass)

Holmstead et al.
1978b

32 Etofenprox

Methanol
5.8 days

RPR 3000 UV lamp


(290320 nm,
Pyrex glass)

Class et al. 1989

59 Diflubenzuron

Methanol
N.R.

RUL 3000 lamp


Ruzo et al. 1974
(>285 nm, borosilicate glass)

63 Xylylcarb

Ethanol
22.6 hr

UV light (>265 nm)

64 Trimethacarb

Cyclohexane
N.R.

1 kW Xe-Hg lamp
Addison et al.
(>300 nm, Corning
1974
0-54 filter)

65 Propoxur

Organic solvents
1239 hr

150 W Hg lamp
Schwack and Kopf
(>280 nm, WG295
1992
filter)

66 Aminocarb

Cyclohexane
N.R.

1 kW Xe-Hg lamp
Addison et al.
(>300 nm, Corning
1974
0-54 filter)
UV light (>265 nm) Kumar et al. 1974

Ethanol
16.4 hr

Kumar et al. 1974

69 Ethiofencarb

Organic solvents
1.35.5 hr

150 W Hg lamp
Kopf and Schwack
(>280 nm, WG295
1995
filter)

78 Pirimicarb

Organic solvents
60140 min

150 W Hg lamp
Schwack and Kopf
(>280 nm, WG295
1993
filter)

96 Chlorsulfuron

Methanol
6.3 hr

20W low-pressure
Hg lamp

Yang et al. 1999

102

T. Katagi

Table 8. (Continued).

No.

Pesticide

Medium
DT50

Light source
(wavelength or
season, UV filter)

Reference

98 Metsulfuron
methyl

Methanol
1.8 hr

20W low-pressure
Hg lamp

Yang et al. 1999

99 Tribenuron
methyl

2-Propanol
1.7 hr

125W mediumpressure Hg lamp

Bhattacharjee and
Dureja 1999

102 Chlorimuronethyl

Hexane
55.8 min

125W Mediumpressure Hg lamp


(Pyrex glass)

Choudhury and
Dureja 1997b

106 Captan

Organic solvents
37420 min

150W Hg lamp
(>280nm, WG295
filter)

Schwack and
Floer-Muller
1990

107 Folpet

Cyclohexene
N.R.

150W Hg lamp
Schwack 1990
(WG295, 305, 320,
335 & 345 filters)

108 Procymidone

Organic solvents
N.R.

150W Hg lamp
(>280nm, WG295
filter)

Schwack et al.
1995b

109 Iprodione

Organic solvents
N.R.

150W Hg lamp
(>280nm, WG295
filter)

Schwack et al.
1995a

110 Vinclozolin

Organic solvents
N.R.

150W Hg lamp
(>280nm, WG295
filter)

Schwack et al.
1995c

125 Endosulfan

Hexane
N.R.

High-pressure Hg
lamp (>300 nm)

Dureja and Mukerjee 1982

130 DDT

Methyl oleate
N.R.

150W Hg lamp
(>280 nm)

Schwack 1988

134 Methoxychlor

Methyl oleate
N.R.

150W Hg lamp
(>280 nm)

Schwack 1988

135 Parathion

12-Hydroxystearate/TL
N.R.
2-Propanol
N.R.

UV-B Fluorescent
Schwack et al.
sunlamp (max, 315
1994
nm)
1kW Tungstenhalogen lamp
(>280 nm)
150W Hg lamp
Schwack 1987
(>280 nm, WG295
filter)

Cyclohexene
N.R.
138 Fenitrothion

Hexane
85 min

Low-pressure u.v.
Pen Ray lamp
(253.7 nm)

Greenhalgh and
Marshall 1976

Photodegradation of Pesticides

103

Table 8. (Continued).

No.

Pesticide

Medium
DT50

Light source
(wavelength or
season, UV filter)

Reference

Methanol
120 min

Low-pressure u.v.
Pen Ray lamp
(253.7 nm)

141 Iodofenphos

Hexane
N.R.

1 kW Metal halide
lamp (Applied
Photophysics
9500)

Walia et al. 1989b

145 Chlorpyrifos

Hexane
N.R.

High-pressure Hg
lamp

Walia et al. 1988

154 Tetrachlovinphos Hexane


4.5 hr

Medium-pressure Hg Dureja et al. 1987b


lamp

170 Phosalone

Hexane
N.R.

125 W High-pressure Walia et al. 1989a


Hg lamp (254360
nm)

188 Triadimefon

Methanol
N.R.

192 Hexaconazole

Hexane
23.1 hr

125 W Hg lamp,
Pyrex

Santoro et al. 2000

193 Penconazole

Organic solvents
523 hr

Tungsten halogen
lamp (WG305 or
WG320 filter)

Schwack and Hartmann 1994

195 Fluotrimazole

Methanol
N.R.

100 W mediumClark et al. 1983


pressure Hg lamp
(borosilicate glass)

211 Anilazine

Methyl oleate
N.R.
Cyclohexene
8.0 or 15.0 min

Metal halogen lamp


(WG295 filter)
Metal halogen lamp
(WG295 or WG
320 filter)

220 Fipronil

Methanol
N.R.

UV light (max. 300


Hainzl and Casida
nm, cutoff of 280
1996
290 nm)

228 Chinomethionat

Unsat. fatty acids/ Fluorescent black


TL
light
N.R.

400 W MediumClark et al. 1978


pressure Hg lamp
(borosilicate glass)
Hexane, methanol 125 W mediumNag and Dureja
2.52.8 hr
pressure Hg lamp
1997

N.R.: not reported; TL, as a thin layer.

Breithaupt and
Schwack 2000

Nutahara and
Murai 1984

104

T. Katagi

Table 9. Photodegradation of pesticides in plant waxes and cuticles.

No.

Pesticide

Wax (W),
Cuticle (C)
DT50

Light source
(wavelength, filter)

Reference

66 Aminocarb

Nectarine fruits W. 125 W Hg lamp (>290


59 min
nm, borosilicate glass)

Pirisi et al. 2001

68 Methiocarb

Nectarine fruits W. 125 W Hg lamp (>290


436 min
nm, borosilicate glass)

Pirisi et al. 2001

78 Primicarb

Nectarine fruits W.
222 min
Fruits W.
35449 min
Fruits W.
15331 min
117 Chlorothalonil Tomato fruits C.
N.R.

125 W Hg lamp (>290


nm, borosilicate glass)
125 W Hg lamp (>290
nm, borosilicate glass)
Sunlight
(MayJune, 39 N)
Simulated sunlight
(Suntest CPS+)

129 2,3,7,8-TCDD

Laurel cherry W.
49 hr

Sunlight and 300 W


high-pressure Hg
lamps

Schuler et al.
1998

135 Parathion

Fruits C.
2.113.5 hr

UV-fluorescent sunlamp
(max, 315 nm)

Schynowski and
Schwack 1996

143 Fenthion

Nectarine fruits W. 125 W Hg lamp (>290


204 min
nm, borosilicate glass)

Pirisi et al. 2001

Fruits W.
2.411.9 hr

Cabras et al.
1997b

N.R., not reported.

Sunlight

Pirisi et al. 2001


Pirisi et al. 1998

Jahn et al. 1999

#
2,4-D

Mecoprop

Triclopyr

Permethrin

Tefluthrin

Cypermethrin

17

18

19

Pesticide

Light source (wavelength or


season, filter), temp (C)

Four soils (-/4.4 7.6/-), 30 m, 33 g cm2


Acetone soln.
450W Medium-pressure
3.3 7.8 d/N.R.
Hg lamp (Pyrex glass)
Three soils (1.4 2.1/-/-), 1 mm, 0.125 g cm2
Methanol soln.
Sunlight (Oct.), outdoors
10 15 d/N.R.
Silty clay loam soils, field (1 5/-/-), -, 34 g cm2
Formulation
Sunlight
75 81 d/N.R.
Dunkirk silt loam soil (2.6/6.0/-), 0.25 mm, 0.02 g cm2
14C, N.R.
Sunlight
N.R.
Loam soil (5.0/6.5/-), -, 600 g ha1
14C, N.R.
Xe lamp (4.5 hr d1)
>31 d/N.R. 25C
Three Japanese soils (2 15/5 6/1 12), 0.5mm, 1.1g cm2
14C, diethyl ether soln.
Sunlight (Aug.), outdoors
0.6 1.9 d/>7 d
Sandy loam soil (1.8/6.9/-), -, 14C, N.R.
Sunlight
56 d/76 100 d

Application mediumb)
DT50 (light/dark)

Soil properties and application ratea)

Table 10. Photodegradation of pesticides in and on soil (clay) thin-layer surfaces.

EPA FIFRA 1999

Takahashi et al. 1985a

EPA FIFRA 1999

Holmstead et al. 1978a

Norris et al. 1987

Romero et al. 1998

Hautala 1978

Reference

Photodegradation of Pesticides
105

#
Cyfluthrin

Cyhalothrin

Deltamethrin

Fenpropathrin

Fenvalerate

Esfenvalerate

20

21

22

24

27

28

Pesticide

Table 10. (Continued).

Light source (wavelength or


season, filter), temp (C)

Sandy loam soil (2.2/5.4/-), -, N.R.


Sunlight
5.3 d/97.9 d
Richvale, MSF and Kracaws types (N.R.), -, 50 g cm2
Hexane soln.
Sunlight (Jun.)
N.R.
Loam soil (5.0/6.5/-), -, 40 g ha1
14C, N.R.
Artificial light, 25C
>166 hr/N.R.
N.R.
14C, N.R.
Xe lamp, continuous
9 d/10 11 d
Kodaira light clay soil (15/5.5/11), 0.5 mm, 1.1 g cm2
14C, diethyl ether soln.
Sunlight (Sept.), outdoors
1 d/>2 wk
Three Japanese soils (2 8/5.7 6.6/4 17), 1 mm, 10 ppm
14C, acetonitrile soln.
500W Xe lamp (>290 nm,
3 51 d/4 160 d
Pyrex glass), 25C
Three Japanese soils (2 15/5 6/4 12), 0.5mm, 0.6g cm2
14C, diethyl ether
Sunlight (Sept.), outdoors
1.8 18 d/>20 d
Noichi sandy clay loam soil (1.4/5.7/3.7), 2 mm, 10 ppm
14C, 1,2-dichloroethane
500W Xe lamp (>300 nm,
100 d/138 d
Pyrex glass), 25C
14C,

Application mediumb)
DT50 (light/dark)

Soil properties and application ratea)

Katagi 1991

Mikami et al. 1980

Katagi 1993b

Takahashi et al. 1985b

EPA FIFRA 1999

EPA FIFRA 1999

Ruzo et al. 1987

EPA FIFRA 1999

Reference

106
T. Katagi

#
Flucythrinate

Fluvalinate

Propachlor

Alachlor

Metolachlor

Metalaxyl

Niclosamide

29

30

33

34

35

37

40

Pesticide

Table 10. (Continued).

Light source (wavelength or


season, filter), temp (C)

Four soils (0.3 0.5/6 8.4/-), 5 mm, 7 mg cm2


Diethyl ether soln.
Sunlight (at 300 400 nm,
1 2 d/>5 d
1.6mW)
Sterilized sandy loam soil (N.R.), 3 5 mm, 1.3 g cm2
14C, N.R.
Sunlight, outdoors
1 d/N.R.
Three soils (0.9 3.5/7.0 7.5/-), 1 mm, 5 20 ppm
Methanol soln.
Sunlight (Jul.)
14 32 d/N.R.
Three Taiwan soils (1.5 2.6/5.0 6.4/-), -, 100 ppm
Acetone soln.
Sunlight (Mar. Apr.),
>8 hr/N.R.
30 35C
Silt loam soil (N.R.), -, 5.15 hg ha1
N.R.
Sunlight, 50 55C
8 d/N.R.
Four soils (0.2 4.6/5 7/60%MWHC), 5 mm, 0.5 g cm2
Aqueous soln.
Xenon lamp (>285 nm) ,25C
8 21 d/36 73 d
Two soils (0.86 1.0/6.1 7.4/-) , -, 500 ppm
N.R.
Sunlight
7.7 d/N.R.
Loamy sand soil (-/5.4/dry or 75%FMC), 2mm, 2.5ppm
14C, acetonitrile soln.
Xe lamp (>290 nm,
7 14 d/18 d
Heraeus suntest), 25C

Application mediumb)
DT50 (light/dark)

Soil properties and application ratea)

Frank et al. 2002


Graebing et al. 2002

Saha and Sukul 1997

Sukul and Spiteller 2001

Chesters et al. 1989

Fang 1977

Konstantinou et al. 2001

Quinstad and Staiger 1984

Dureja and Chattopadhyay 1995

Reference

Photodegradation of Pesticides
107

#
Carboxin

Napropamide

Florasulam

Monuron

Diuron

Linuron

Isoproturon

Thidiazuron

42

47

48

52

53

54

56

58

Pesticide

Table 10. (Continued).

Light source (wavelength or


season, filter), temp (C)

Scheyern soil (N.R.), 2 mm, 5 ppm


N.R.
Xe lamp (>290 nm,
29 hr/N.R.
Heraeus suntest),18C
Two soils (0.3 1.7/7.3 7.9/-), -, CH2Cl2 soln.
Sunlight
3 7 d/N.R.
Caltin silt loam soil (2.9/6.8/dry), -, 0.13 0.32 ppm
14C,
N.R. Sunlight (May Jun.)
30 d/79 d
Sand/montmorillonite and kaolinite clays, -, Diethyl ether soln.
GL20 fluorescent lamp
N.R.
(>300 nm, glass filter)
Sand/montmorillonite and kaolinite clays, -, Diethyl ether soln.
GL20 fluorescent lamp
20 100 hr/N.R.
(>300 nm, glass filter)
Silt loam soil (N.R.), -, 14C, N.R.
Xe lamp (Pyrex glass),
>15 d/N.R.
25C
Sandy loam soil (0.35/7.2/-), -, 2.5 mg/g of soil
N.R.
Low-pressure Hg lamp
N.R.
Speyer soil 2.3 (1.2/5.5/1.5), 0.5 mm, 1.5 g cm2
14C, methanol soln.
2.5kW Xe lamp (>290 nm,
0.5 hr/51 hr
WG295+Duran filter), <30C

Application mediumb)
DT50 (light/dark)

Soil properties and application ratea)

Klehr et al. 1983

Kulshrestha and Mukerjee 1986

EPA OPPTS 1995d

Jirkovsky et al. 1997

Jirkovsky et al. 1997

Krieger et al. 2000

Donaldson and Miller 1996

Murthy et al. 1998

Reference

108
T. Katagi

#
Diflubenzuron

Propoxur

Methiocarb

Carbofuran

Benomyl

Asulam

Desmedipham

Phenmedipham

59

65

68

72

81

82

83

84

Pesticide

Table 10. (Continued).

N.R.

Light source (wavelength or


season, filter), temp (C)

N.R.
Artificial light
11.3 d/3.7 d
Two soils (1.0 1.5/5.3 7.1/-), -, 0.4 0.6 ppm
14C, methanol
Medium-pressure Hg lamp
>1 d/ 200d
(>290 nm, TQ150 filter)
Four soils (4.5 7.5/5.8 6.3/dry), -, CH2Cl2soln.
Sunlight (Jun. Aug.,
7 14 d/N.R.
Kimax glass), outdoors
Orange glove soil (N.R.), -, 44.8 g cm2
2.5EC as (49)
Sunlight
N.R.
Silt loam soil (N.R.), -, 1 lb acre1
14C, N.R.
Sunlight at 25C
< 4 d/N.R.
Sandy loam soil (N.R.), -, 14C, N.R.
Xe lamp, 25 28C.
1.5 hr/83 hr
Sandy loam soil (N.R.), -, 14C, N.R.
Xe lamp (3-fold irradiation
110 160 hr/>500 hr
of summer noon sunlight)
Metapunto soil (4.8/7.4/-), -, Acetonitrile soln.
Xe lamp
12 d/15 d
(Heraeus suntest)

14C,

Application mediumb)
DT50 (light/dark)

Soil properties and application ratea)

Schafmeier et al. 1998

EPA OPPTS 1996b

EPA OPPTS 1995a

EPA OPPTS 2001

Nigg et al. 1984

Gohre and Miller 1986

EPA OPPTS 1997b

EPA OPPTS 1997a

Reference

Photodegradation of Pesticides
109

#
Benthiocarb

Butyrate

Carbosulfan

Benfuracarb

Molinate

Thiophanatemethyl
Maneb

Chlorsulfuron

86

87

88

89

90

92

94

96

Pesticide

Table 10. (Continued).

Light source (wavelength or


season, filter), temp (C)

Rice-growing area clay soil (1.95/4.6/-), -,acetone soln.


Sunlight
N.R.
(3.5 4.5 mW cm2)
Loam soil (N.R.), -, 14C, N.R.
Sunlight, outdoors, 25C
N.R.
Orange glove soil (N.R.), -, 44.8 g cm2
2.5 EC
Sunlight
5.9 d/N.R
Sandy loam soil (N.R.), -, 125 g cm2
N.R.
Low-pressure Hg lamp
N.R.
(254 nm)
Three soils (0.9 3.5/7.0 7.5/-), 1 mm, 5 20 ppm
Methanol soln.
Sunlight (Jul.)
13 34 d/N.R.
Sandy loam soil (1.7/7.4/-), -, 14C, N.R.
Sunlight
2.9 5.5 d/10 19 d
Delaware soil (N.R.), -, 22.4 g cm2
14C, aqueous soln.
Sunlight (spring)
<1 wk/N.R.
Nora silty clay loam soil (2/8.0/-), 1 mm, 1.7 g cm2
14C, pH 7 buffer soln.
Xe lamp (>290 nm), 25 C
50 d/130 d
14C,

Application mediumb)
DT50 (light/dark)

Soil properties and application ratea)

Strek 1998

Rhodes 1977

EPA OPPTS 2001

Konstantinou et al. 2001

Dureja et al. 1990

Nigg et al. 1984

EPA OPPTS 1993

Cheng and Hwang 1996

Reference

110
T. Katagi

#
Tribenuronmethyl
Triasulfuron

Chlorimuronethyl
Rimsulfuron

Thifensulfuronmethyl
Captan

Iprodione

Famoxadone

99

100

102

103

105

106

109

111

Pesticide

Table 10. (Continued).

Light source (wavelength or


season, filter), temp (C)

Alluvial (1.2/7.5/-), 2 mm, acetone soln.


Sunlight (May) or u.v.
11 d or 11 hr/N.R.
light (max. 254 nm)
Two soils (0.9 2.3/7.0 7.1/dry) , 1 mm, 6 7 ppm
Methanol soln.
1.1kW Xe lamp (Suntest
6 15 hr/26 198 hr
CPS+, >300 nm), 20 C
Three soils (0.7 1.2/6.2 8.1/-), 2 mm, 50 ppm
N.R.
Sunlight (Apr. May),
11 21 d/N.R.
30 35 C
Sassafras sandy loam soil (1.0/6.3/-), 1mm, 0.5g cm2
14C, acetonitrile soln.
Sunlight (Jun. Jul.), 25C
11 d/11 d
Two soils (0.9 2.3/7.0 7.1/dry) , 1 mm, 6 7 ppm
Methanol soln.
1.1kW Xe lamp (Suntest
7 hr/24 34 hr
CPS+, >300 nm), 20 C
Sandy loam soil (-/-/moist), -, 14C, N.R.
Sunlight
5 15 d/10 21 d
Sandy loam soil (N.R.), -, 14C, N.R.
Xe lamp
7 14 d/14 21 d
N.R.
N.R.
N.R.
12 d/28 d
14C,

Application mediumb)
DT50 (light/dark)

Soil properties and application ratea)

Jernberg and Lee 1999

EPA OPPTS 1998e

EPA OPPTS 1999b

Albanis et al. 2002

Schneiders et al. 1993

Choudhury and Dureja1997a

Albanis et al. 2002

Bhattacharjee and Dureja 2002

Reference

Photodegradation of Pesticides
111

#
Dicamba

Chloramben

Pentachlorophenol
Aldrin

Dieldrin

DDT

Dicofol

118

119

121

122

123

130

133

Pesticide

Table 10. (Continued).

Light source (wavelength or


season, filter), temp (C)

Laponite,-,N.R.
125W high-pressure Hg lamp
N.R.
(>290 nm, 2.2-cm water filter)
Pennsylvania loam soil (2.0/56/-), 2 mm, 3 ppm
14C, acetonitrile
Xenon lamp (Heraeus
29 109d/2 500d
Suntest CPS), 25C
Three soils (0.3 1.5/4.4 7.7/dry), 0.08 mm, 1000 ppm
Hexane soln.
300W medium-pressure
20 60 min/N.R.
Hg lamp (water filter)
UK and German soils (2 3.5/7.4 8.1/-), -, 30 g cm2
14C, EC formulation
Sunlight (April), outdoors
N.R.
Japanese soils (N.R.), -,Formulation
Sunlight, filed
N.R.
Clay loam soil (2.5/7.5/field), -, 14C, N.R.
Sunlight, outdoors
55 d/N.R.
Silt loam soil (N.R.), -,14C, N.R.
Artificial light
21 30 d/N.R.

Application mediumb)
DT50 (light/dark)

Soil properties and application ratea)

EPA OPPTS 1998d

Zayed et al. 1994

Suzuki and Yamamoto 1974

Klein et al. 1973

Liu et al. 2002

Misra et al. 1997

Aguer et al. 2000

Reference

112
T. Katagi

#
Parathionmethyl

Cyanophos

Fenitrothion

Bromophos

Iodofenphos

Tolclofosmethyl

136

137

138

140

141

142

Pesticide

Table 10. (Continued).

Light source (wavelength or


season, filter), temp (C)

Orthic livisol (1.9/7.2/-), 0.1 mm, Methanol soln.


Xe lamp in Suntest CPS+
1 d/>1 d
20.5C
Sandy loam soil (N.R.), -, >14 g cm2
14C, N.R.
Sunlight (July), outdoors
61 d/106 d
ca. 25C
Two Japanese soils (3 19/5.2 6.4/5 12), 0.5 mm, 14C, N.R.
Sunlight (Sept. Oct.),
2 d/>12 d
outdoors
Two Japanese soils (2.5 19/-/5 12), 50 m, 10 g cm2
14C, CHCl soln.
Sunlight (Jun.), outdoors
3
1 d/>12 d
Speyer standard soil 2.2 (2.6/5.8/12), -, 180 ppm
Methanol soln.
Daylight + UV-A & -B lamps
48 d/80 d
(>290 nm), 25C
Finely powdered soil (N.R.), -, Acetone soln.
1 kW metal halide lamp
N.R.
(Applied Photophysics 9500)
Speyer standard soil 2.22 (2.6/5.8/12), -, 180 ppm
Methanol soln.
Daylight + UV-A & -B lamps
71 d/85 d
(>290 nm), 25C
Four Japanese soils (3 15/5 7/5 16) , 50 m, 7 g cm2
14C, CHCl soln.
Sunlight (May Jun.),
3
1 2 d/2 >15 d
outdoors

Application mediumb)
DT50 (light/dark)

Soil properties and application ratea)

Mikami et al. 1984b

Allmaier et al. 1984

Walia et al. 1989b


Allmaier and Schmid 1985

Allmaier et al. 1984


Allmaier and Schmid 1985

Mikami et al. 1985b

Mikami et al. 1976

EPA OPPTS 1999e

Kromer et al. 1999

Reference

Photodegradation of Pesticides
113

#
Fenthion

Diazinon

Chlorpyrifos

Quinalphos

Tetrachlovinphos

Monocrotophos

143

144

145

148

154

156

Pesticide

Table 10. (Continued).

Light source (wavelength or


season, filter), temp (C)

Five U.S. soils (0.8 6.3/4.5 7.5/-), -, 50 200 ppm


CH2Cl2 soln.
Sunlight (May Aug.),
N.R.
outdoors
Switzerland silty loam (3.6/6.1/dry or 12%), -, 10 ppm
14C, N.R.
Xe lamp (>290 nm, u.v. filter
<1 d/<1 d
+ IR reflecting glass), 45 C
Sandy loam soil (N.R.), -, 14C, N.R.
Sunlight
20 hr/ 14.7 d
Finely powdered soil (N.R.), 2 mm, 0.5% acetone soln.
Low-pressure Hg lamp
17.3 d/N.R.
(254 nm)
Four Indian soils (-/5.6 8.4/-), 50 m, 7 g cm2
Chloroform soln.
Sunlight (May Jun.)
2 5 d/14 >25 d
East Anglia medium loam (-/8.0/19),-, 13 ppm
14C, acetone soln.
Subdued daylight
4 5 d/N.R.
Sandy loam soil (0.35/7.2/ -), -, 2.5 mg/g of soil
N.R.
Medium-pressure Hg lamp
N.R./>10 d
Sterilized sandy loam (1.3/6.4/-) soil, <1 mm, 5 ppm
14C, methanol soln.
Sunlight (Jul. Aug.)
3 d/30 d

Application mediumb)
DT50 (light/dark)

Soil properties and application ratea)

Lee et al. 1989

Dureja et al. 1987b

Beynon and Wright 1969

Dureja et al. 1988

Walia et al. 1988

EPA OPPTS 2000

Burkhard and Guth 1979

Gohre and Miller 1986

Reference

114
T. Katagi

#
Dicrotophos

Profenofos

Phorate

Disulfoton

Bensulide

Phenthoate

Azinphosmethyl

157

160

162

163

164

168

169

Pesticide

Table 10. (Continued).

Light source (wavelength or


season, filter), temp (C)

Four U.S. soils (-/5.7 6.7/-), <1 mm, 25 ppm


acetone soln.
Fluorescent lamps
N.R.
(GEP40BL, BLB), 35C
Switzerland silty loam (3.6/6.1/dry or 12%), -, 10 ppm
14C, N.R.
Xe lamp (>290 nm, u.v. filter
<1 d/ 1 d
+ IR reflecting glass), 45 C
Plano silt loam soil (N.R.), field, 112 g cm2
Formulation
Sunlight (May), field
<1 wk/N.R.
Sandy loam soil (N.R.), -, 14C, N.R.
Sunlight
2.4 d/N.R.
Four soils (4.5 7.5/0.8 6.3/dry), -, Dichloromethane soln.
Sunlight (Aug. or Oct.,
3 d/>5 d
Kimax glass), outdoors
Sorento loam soil (N.R.), -,14C, N.R.
Xe lamp
90 d/N.R.
Two Japanese soils (3 19/5 6/5 12), 0.5 mm, 10 g cm2
14C, diethyl ether soln.
Sunlight (Sept. Oct.),
3 d/>6 d
outdoors
Sandy loam soil (-/5.1/-), -, 14C, N.R.
Sunlight (April)
180 d/N.R.
14C,

Application mediumb)
DT50 (light/dark)

Soil properties and application ratea)

EPA OPPTS 1998a

Mikami et al. 1977b

EPA OPPTS 1999a

Gohre and Miller 1986

EPA OPPTS 1999d

Lichtenstein et al. 1973

Burkhard and Guth 1979

Lee et al. 1989

Reference

Photodegradation of Pesticides
115

#
Phosalone

Methidathion

Isofenphos

Ditalimfos

Cyanofenphos

Dioxabenzofos

Ethion

Atrazine

170

171

175

178

180

183

184

185

Pesticide

Table 10. (Continued).

Light source (wavelength or


season, filter), temp (C)

Finely powdered soil (N.R.), -, Acetone soln.


1kW high-pressure metal5 d/>15 d
halide lamp (>300 nm)
Switzerland silty loam (3.6/6.1/dry or 12%), -, 10 ppm
14C, N.R.
Xe lamp (>290 nm, u.v. filter
<1 d/<1 d
+ IR reflecting glass), 45 C
Sandy loam soil (0.35/7.2/-), -, 125 g cm2
N.R.
Low-pressure Hg lamp
N.R.
Speyer standard soil 2.2 (2.6/5.8/12), -, 310 ppm
Methanol soln.
Daylight + UV-A & -B
18 d/N.R.
lamps (>290 nm), 25C
Two Japanese soils (3 19/5.2 6.4/5 12), 0.5 mm, 14C, N.R.
Sunlight (Sept. Oct.),
2 d/>12 d
outdoors
Two Japanese soils (3 19/5 6/5 12), 0.5 mm, 10 g cm2
14C, diethyl ether soln.
Sunlight (Sept. Oct.),
<1 d/>6 d
outdoors
Sterilized sandy loam soil (N.R.), -, 8.6 ppm
14C, N.R.
Sunlight
61.6 d/175.9 d
Eight soils (0.9 1.6/7.8 8.0/various), 0.2 0.5 mm, Methanol soln.
1kW medium-pressure
4 8 min/N.R.
Hg lamp

Application mediumb)
DT50 (light/dark)

Soil properties and application ratea)

Gong et al. 2001

EPA OPPTS 1995c

Mikami et al. 1977b

Mikami et al. 1976

Allmaier et al. 1984

Dureja et al. 1989

Burkhard and Guth 1979

Walia et al. 1989a

Reference

116
T. Katagi

#
Triadimefon

Propiconazole

Amitrole

Terbacil

Bentazon

Metribuzin

188

191

196

197

200

201

Pesticide

Table 10. (Continued).

Light source (wavelength or


season, filter), temp (C)

Two soils (0.4 0.8/5.3 7.4/-), 2 mm, 31.8 g cm2


Acetone soln.
u.v. light (254 nm) or
9.5 22 hr/N.R.
sunlight (Sept.)
Scheyern soil (N.R.), 2 mm, 5 ppm
N.R.
Xe lamp (>290 nm,
166 hr/N.R.
Heraeus suntest), 18C
Sandy loam soil (N.R.), 3.5 mm, 0.3g cm2
Acetone soln.
Sunlight
12 d/>26 d
Sandy loam soil (N.R.), -, 14C, N.R.
Xe lamp, 25C
<73 d/N.R.
Drummer silty clay loam soil (N.R.), -, 1.2 lb acre1
14C, N.R.
Xe lamp, 25C
61 d/N.R.
Montcalm sandy loam soil (N.R.), 0.5 mm, 100 g cm2
14C, CH Cl soln.
RPR u.v. light (300 &
2 2
>5 d/N.R.
350 nm)
Sandy loam soil (N.R.), -, 14C, N.R.
Sunlight, outdoors, 31C
2.5 d/N./R.

Application mediumb)
DT50 (light/dark)

Soil properties and application ratea)

EPA OPPTS 1998g

Nilles and Zabik 1975

EPA OPPTS 1998h

EPA OPPTS 1996a

Dureja et al. 1987a

Murthy et al. 1998

Nag and Dureja 1996

Reference

Photodegradation of Pesticides
117

#
Norflurazon

Fipronil

DTP

Trifluralin

Butralin

Oryzalin

214

220

221

232

234

235

Pesticide

Table 10. (Continued).

Light source (wavelength or


season, filter), temp (C)

Rome loam soil (2.0/6.2/-), -, 0.6 g cm2


N.R.
Sunlight, outdoors, 25 70C
8 d/N.R.
Northern Senegal soil (0.3/7.6/-), field, 0.05 0.1 g cm2
Formulation
Sunlight, field
45 d/N.R.
Two Niger soils (0.1 0.3/5.3 5.8/-) , -, 0.08 g cm2
Formulation
Sunlight (Sept. Oct.), field
1.5 d/N.R.
Banizoumbou soil (6.5/8.3/dry) soils, -, 2.5 ppm
Methanol soln.
1.8 kW Xe lamp (Suntest,
69 d/N.R.
>290 nm, u.v.filter)
Two Japanese soils (3 9/5.6 5.8/-), 0.5 mm, 3.8 g cm2
14C, aqueous soln.
Sunlight (May Jun.)
31 34 d/N.R.
Sandy loam soil (N.R.), -, 14C, N.R.
N.R.
41 d/66 d
Sandy loam soil (N.R.), -, 374 ppm
14C, N.R.
Sunlight
99.6 d/112.7 d
Sandy loam soil (N.R.), -, 14C, N.R.
Xe lamp
22.4 hr/N.R.

Application mediumb)
DT50 (light/dark)

Soil properties and application ratea)

EPA OPPTS 1994

EPA OPPTS 1998b

EPA OPPTS 1996c

Yamaoka et al. 1988

Bobe et al. 1998b

Bobe et al. 1998a

Fenet et al. 2001

Schroeder Kvien and Banks 1985

Reference

118
T. Katagi

Fluchloralin

Ethalfluralin

Pendimethalin

Prochloraz

236

237

238

240

Pesticide

Light source (wavelength or


season, filter), temp (C)

Montcalm sandy loam soil (N.R.), 0.5 mm, 100g cm2


CH2Cl2 soln.
RPR u.v. light (300 &
N.R.
350 nm)
Sandy loam soil (N.R.), -, 14C, N.R.
N.R.
14.2 d/N.R.
Kalyani alluvial soil (1.7/7.2/dry), 2mm, 28g/g of soil
Hexane soln.
Sunlight
N.R.
New Delhi sandy loam soil (3.6/6.1/-), 50 m, 7g cm2
Methanol soln.
Low-pressure Hg lamp
N.R.
(254 nm)
Two soils (1.0 1.5/5.3 7.1/-), -, 0.4 0.6 ppm
14C, methanol
Medium-pressure Hg lamp
>1 d/ 200d
(>290 nm, TQ150 filter)
14C,

Application mediumb)
DT50 (light/dark)

Soil properties and application ratea)

Hollrigl-Rosta et al. 1999

Dureja and Walia 1989

Hadler et al. 1989

EPA OPPTS 1995b

Nilles and Zabik 1974

Reference

aSoil properties and application rate are listed in the following order as boldface: soil name (organic matter content %/pH/ moisture content %), soil layer thickness,
application rate.
bNonradiolabeled pesticide when unspecified.
- or N.R.: not reported. MHWC, maximum water holding capacity; FMC, field moisture content

Table 10. (Continued).

Photodegradation of Pesticides
119

Phenothrin
Cypermethrin
Deltamethrin

Fenpropathrin
Fenvalerate

Flucythrinate

19
22

24
27

29

2,4-D

Pesticide

13

No.

French bean leaf


3d

Kidney bean leaf


14 d
Spring wheat leaf
24 d
Cotton leaf
8d

Mandarin orange leaf


4d

Cotton leaf
1.1 wk
Bean leaf
4 d
Potato leaf
12 d

Cotton & bean leaf


24 d

Bean & rice leaf


<1 d

Corn leaf
14.6 hr

Plant species
DT50

Cole et al. 1982


Ruzo and Casida 1979

methanol soln.
Sunlight (Nov.), greenhouse

14C,

diethyl ether soln.


Sunlight (JulySept.), outdoors

14C,

Methanol soln.
Sunlight (Apr.), greenhouse

Chattopadhyaya and Dureja 1991

Holmstead et al. 1978b

Lee et al. 1988

Ohkawa et al. 1980

14C,

methanol soln.
Sunlight (Oct.), greenhouse
14C, 0.1% EC
Sunlight (Sept.Dec.), outdoors
2.4 lb/gal EC formulation
Sunlight (mid-July)

Takahashi et al. 1985b

methanol soln.
Sunlight (Sept.), greenhouse

14C,

Maguire 1990

Cole et al. 1982

Nambu et al. 1980

14C,

diethyl ether soln.


Sunlight, greenhouse
14C, diethyl ether soln.
Sunlight (JulySept.), outdoors
2.5% EC
Sunlight (summer), field

Venkatesh and Harrison 1999

References

0.25% oxysorbic surfactant


Fluorescent lamps

14C,

Application medium
Light source (season), conditions

Table 11. Surface degradation of pesticides in foliar or topical application.

120
T. Katagi

Flutolanil
Carboxin
Propyzamide
Diafenthiuron

Diflubenzuron

Metolcarb
Fenobucarb
Xylylcarb
Trimethacarb

42
45
57

59

61
62
63
64

Pesticide

39

No.

Table 11. (Continued).

Pinto bean leaf


N.R.

Bean leaf
<1 d

Rice leaf
N.R.

Bean leaf
<1 d

Cotton leaf
31.9 d
Conifer pine needle
23 wk

Chinese cabbage leaf


4d
Cotton leaf
<3 hr

Alfalfa leaf
63.8 d

Bean leaf
<10 hr

Cucumber leaf
29.3 d

Plant species
DT50

Ogawa et al. 1976


Ohkawa et al. 1974a
Slade and Casida 1970

14C,

methanol-water (1/3) soln.


Sunlight, greenhouse

14C,

methanol soln.
Sunlight, greenhouse
14C,

ethanol soln.
Sunlight (JuneJuly), outdoors

Ohkawa et al. 1974a

Rodriguez et al. 2001

Bull and Ivie 1976

Drabek et al. 1992

methanol soln.
Sunlight, greenhouse

14C,

25% WP
Sunlight (summer), outdoors
45% oil formulation
Sunlight (Aug.Oct.), field

14C,

25 EC
Sunlight, field
SC400 formulation
Sunlight, field

Keum et al. 2002

Yih and Swithenbank 1971

14C,

acetone soln.
Sunlight, field

Buchenauer 1975

Uchida et al. 1983

References

Suspension
Sunlight

diethyl ether soln.


Sunlight, greenhouse

14C,

Application medium
Light source (season), conditions

Photodegradation of Pesticides
121

Propoxur
Aminocarb
Mexacarbate
Methiocarb
Carbaryl
Carbofuran
Oxamyl
Pirimicarb
Benomyl
Fenothiocarb
Carbosulfan

66
67
68
71
72
74
78
81
85
88

Pesticide

65

No.

Table 11. (Continued).

Valencia orange
36 d

Mandarin orange
1.612 d

Pinto bean leaf


N.R.

Lettuce leaf
2 d

Tobacco leaf
15 d

Strawberry
13 d

Garden snapbean leaf


2.8 d

Garden snapbean leaf


>3 d

Garden snapbean leaf


23 hr

Garden snapbean leaf


4 hr

Garden snapbean leaf


8 hr

Plant species
DT50

Abdel-Wahab and Casida 1967


Abdel-Wahab and Casida 1967
Abdel-Wahab et al. 1966
Abdel-Wahab et al. 1966

14C,

ethanol soln.
Sunlight (Aug.Sept.), outdoors

3H, 14C;

ethanol soln.
Sunlight (Aug.Sept.), outdoors

14C,

ethanol soln.
Sunlight (Aug.Sept.), outdoors

14C,

Harvey et al. 1978

14C,

Baude et al. 1973


Unai et al. 1986
Clay and Fukuto 1984

14C,

52% WP
Sunlight, greenhouse
14C, 30% EC
Sunlight (Sept.), greenhouse
14C,

EC
Sunlight (summer), field

Cabras et al. 1990

25% liquid formulation


Sunlight (May), field

0.2% Tween 20 aqueous soln.


Plant growth chamber

Archer et al. 1977

43.8% FL
Sunlight, field

ethanol soln.
Sunlight (Aug.Sept.), outdoors

Abdel-Wahab et al. 1966

References

ethanol soln.
Sunlight (Aug.Sept.), outdoors

14C,

Application medium
Light source (season), conditions

122
T. Katagi

Benfuracarb
Thiophanate-methyl

MBC
ETU
Tribenuron methyl
Thifen-sulfuron
Procymidone
Methazole
Aldrin
Dieldrin

92

93
95
99

104

108

112

122

123

Pesticide

89

No.

Table 11. (Continued).

Apple tree leaf


<1 wk

Apple tree leaf


<1 wk

Cotton leaf
<1 d

Bean & cucumber leaf


23 wk

Soybean leaf
420 d

Wheat leaf
N.R.

Tomato leaf
2.1 d

Corn leaf
N.R.

Cotton leaf
N.R.
Apple tree leaf
15 d

Cotton leaf
N.R.

Plant species
DT50

Rhodes 1977

14C,

Brown et al. 1993

aqueous soln. surfactant


Sunlight, greenhouse

50% dispersible powder


Sunlight (June), field

Aldrex 30
Sunlight (June), field

Harrison et al. 1967

Harrison et al. 1967

Dorough et al. 1973

14C,

aqueous soln.
Sunlight (June), outdoors

Mikami et al. 1984a

diethyl ether soln.


Sunlight, greenhouse

14C,

14C,

Bhattacharjee and Dureja 2002

Acetone soln.
Sunlight, greenhouse

80% WP
Sunlight, outdoors

Fleeker and Lacy 1977

Soeda et al. 1972

Buchenauer et al. 1973

Tanaka et al. 1985

References

10 mM HCl in 50% ethanol.


Sunlight (July), outdoors

3H, 14C;

70% WP
Sunlight, outdoors
14C, 70% WP
Sunlight (MayAug.)

acetone-water (1/1)
Sunlight, greenhouse

14C,

Application medium
Light source (season), conditions

Photodegradation of Pesticides
123

Endosulfan

Endrin
DDT
Parathion

Cyanophos
Fenitrothion

Bromophos
Fenthion

126

130

135

137

138

140

143

Pesticide

125

No.

Table 11. (Continued).

Orange fruit, peel


3.94.8 d
Corn leaf
1 d

Tomato leaf
< 8 hr

Apple tree leaf


1.2 d
Bean leaf
<1 d

Kidney bean leaf


0.54 d

Garden bean leaf


1 d
Cotton leaf
N.R.

Apple tree leaf


<12 wk

Apple tree leaf


<1 wk

Apple tree leaf


<1 wk
Cotton leaf
N.R.

Plant species
DT50

Commercial formulation (24.2% a.i.)


Sunlight (Apr.), outdoors
50% EC
Sunlight, field

& 3T, emulsion


400 W daylight lamps in chamber

32P

0.5% emulsion
Sunlight (Oct.Nov.)
14C, ethanol soln.
Sunlight (Nov.), outdoors

14C,

benzene-hexane (4/1) + surfactant


Sunlight, greenhouse

El-Refai and Hopkins 1966

0.2% Triton X-155 + B-1956(1/1)


Fluorescent lamps, greenhouse
14C, methanol soln.
Sunlight, field

Leuch and Bowman 1968

Minelli et al. 1996

Stiasni et al. 1969

Ohkawa et al. 1974b

Hosokawa and Miyamoto 1974

Chiba et al. 1976

Joiner and Baetcke 1973

Harrison et al. 1967

Harrison et al. 1967

Dureja and Mukerjee 1982

Harrison et al. 1967

References

25% EC or 50% dispersible powder


Sunlight (June), field

Endrex 20
Sunlight (June), field

20% concentrate
Sunlight (June), field
0.1% hexane soln.
Sunlight (June)

Application medium
Light source (season), conditions

124
T. Katagi

Chlorpyrifos
Isoxathion
Phoxim

Tetrachlo-vinphos

Monocroto-phos

Dicrotophos
Phosphami-don
Sulprofos

151

152

154

156

157

158

161

Pesticide

145

No.

Table 11. (Continued).

Cotton leaf
1d

Cotton leaf
<1 d

Cotton leaf
<1 d

Cotton leaf
<2 d
Bush apple fruit
N.R.

Apple tree leaf


2 d
Bean leaf
6d

Tomato leaf
1 d
Corn leaf
<1 d

Cabbage leaf
23 d

Soft shield fern leaf


54.3 d

Plant species
DT50

Bull et al. 1967


Ivie and Bull 1976

32P,

aqueous soln.
Sunlight, greenhouse
14C,

50% EC
Sunlight (Aug.), outdoors

Bull and Lindquist 1964

Beynon and Wright 1972

Lindquist and Bull 1967

Dureja et al. 1987b

Beynon and Wright 1969

Bowman and Leuck 1971

aqueous soln.
Laboratory

32P,

aqueous soln.
Sunlight, greenhouse
14C, acetone soln.
Sunlight, greenhouse

32P,

acetone soln.
Sunlight, greenhouse
Acetone soln.
Sunlight

14C,

5% formulation
Sunlight, field
EC
Sunlight, field

Makary et al. 1981

Ando et al. 1975

14C,

0.03% Tween 20 aqueous soln.


Sunlight, greenhouse

Walia et al. 1988

References

0.1% acetone soln.


Solar simulator (1 kW metal-halide lamp)

14C,

Application medium
Light source (season), conditions

Photodegradation of Pesticides
125

Dimethoate

Formothion
Malathion

Phenthoate
Azinphos-methyl
Edifenphos
Acephate

166

167

168

169

173

174

Pesticide

165

No.

Table 11. (Continued).

Cotton leaf
5.9 hr
Tobacco leaf
58 d

Rice leaf
5 d

Bean & corn leaf


1 d

Valencia orange
37 d

Broad bean leaf


N.R.
Bean leaf
2d
Cotton leaf
1.23.8 d

Bean leaf
1.2 d

Corn leaf
N.R.
Bean leaf
4 d
Bean leaf
1.7 d

Plant species
DT50

Ishizuka et al. 1973


Bull 1979

acetone soln.
Sunlight (JuneAug.)

32P,

emulsion
Sunlight, greenhouse
14C,

Yamazaki et al. 1982

Liang and Lichtenstein 1976

14C,

0.1% Triton X-100


Sunlight, field
50% WP
Sunlight, greenhouse

Takade et al. 1976

& 14C, 0.06% emulsion


Sunlight (Mar. & Sept.)

32P

Awad et al. 1967

El-Refai and Hopkins 1972

Mostafa et al. 1974

32P,

acetone soln.
Sunlight, field
Emulsion
Sunlight, field
14C, ULV or 57% EC
Sunlight, greenhouse

Sauer 1972

Lucier and Menzer 1970

Lucier and Menzer 1968

Dauterman et al. 1960

References

20% EC
Artificial light, greenhouse

14C,

emulsion
Sunlight, greenhouse
14C & 32P, aqueous soln.
Sunlight (Apr.), greenhouse
14C, aqueous soln.
Sunlight, greenhouse

32P,

Application medium
Light source (season), conditions

126
T. Katagi

Pesticide
Cyanofenphos
Leptophos
Butonate
Triadimefon
Triadimenol
Fluotrimazole
Lenacil
Dinoseb
Dinobuton
Imidacloprid
Fipronil

No.

180

181

182

188

189

195

199

202

203

215

220

Table 11. (Continued).

Corn leaf
20 hr

Tomato leaf
0.71.4 d

Garden snapbean leaf


0.3 hr

Garden snapbean leaf


0.9 hr

Sugar beet leaf


N.R.

Barley leaf
N.R.

Apple tree leaf


N.R.

Marrow, 1st true leaf


N.R.

Bean leaf
8.3 hr

Cotton leaf
N.R.

Kidney bean leaf


2.25 wk

Plant species
DT50

Matsuo and Casida 1970


Bandal and Casida 1972
Scholz and Reinhard 1999

50% WP
Sunlight, greenhouse

14C,

ethanol soln.
Sunlight (Aug.), greenhouse

14C,

ethanol soln.
Sunlight (Aug.), greenhouse

+ 20SL formulation
Sunlight (51 N), greenhouse
Acetone soln.
Sunlight (Nov.)

Hainzl and Casida 1996


Fenet et al. 2001

Zhang et al. 1999

14C,

14C

Clark et al. 1983

Clark and Watkins 1986

Acetone-water (1/1) soln.


400W Hg lamp, greenhouse

Acetone-water (3/7) soln.


Sunlight (July), outdoors

1 mM, WP
Sunlight, greenhouse

Clark et al. 1978

Derek et al. 1979

32P,

aqueous soln.
Sunlight

Zayed et al. 1978

Chiba et al. 1976

References

50% aqueous acetone


Sunlight

32P,

benzene-hexane (4/1) + surfactant


Sunlight, greenhouse

14C,

Application medium
Light source (season), conditions

Photodegradation of Pesticides
127

Alloxydim
Guazatine
Avermectin B1a
MAB1a
Spinosyn A

224

245

250

251

252

4 plant species
1.616 d

Commercial formulation
Sunlight

Saunders and Bret 1997

Wrzesinski et al. 1996

14C,

Cabbage leaf
N.R.

EC
Sunlight, field

Moye et al. 1990

& 3H, EC formulation


Sunlight (JuneMar.), outdoors

+ 200 ppm nonionic surfactant


Sunlamps, growth chamber

14C

Sato et al. 1985a

aqueous emulsion
Sunlight (Jan.Mar.), outdoors

Celery
59 d

Soeda et al. 1979

14C,
14C,

Campbell and Penner 1985b

References

aqueous emulsion.
Fluorescence light, greenhouse

14C,

Application medium
Light source (season), conditions

Dwarf apple leaf


67 wk

Sugar beet leaf


3d

Navy bean leaf


1 hr

Plant species
DT50

N.R.: not reported.


Formulation: ULV: ultra-low-volume formulation; EC: emulsifiable concentrate; WP: wettable powder; FL: flowable.

Sethoxydim

Pesticide

223

No.

Table 11. (Continued).

128
T. Katagi

Photodegradation of Pesticides

129

Appendices Listing

Appendix Number
1 ................................................................................................................................
2 ................................................................................................................................
3 ................................................................................................................................
4 ................................................................................................................................
5 ................................................................................................................................
6 ................................................................................................................................
7 ................................................................................................................................
8 ................................................................................................................................
9 ................................................................................................................................
10 ................................................................................................................................
11 ................................................................................................................................

Page
136
137
139
140
141
144
145
146
148
151
152

130

T. Katagi
Directory of pesticide chemical structures.
Pesticide
Acephate
Acrinathrin
Alachlor
Aldrin
Allethrin
Alloxydim
Aminocarb
Amitrole
Anilazine
Asulam
Atrazine
Avermectin B1a
Azadirachtin-A
Azinphos-methyl
Azoxystrobin
Benfuracarb
Benomyl
Bensulide
Bentazone
Benthiocarb
Bioallethrin
Bromacil
Bromophos
Buprofezine
Butachlor
Butamifos
Butonate
Butralin
Butylate
Captan
Carbaryl
Carbofuran
Carbosulfan
Carboxin
Cartap
Chinomethionat
Chloramben
Chlordane
Chlordimeform
Chlorimuron
Chlorimuron ethyl
Chlorothalonil
Chlorpropham
Chlorpyrifos

Identification number

Appendix number

174
31
34
122
10
224
66
196
211
82
185
250
253
169
244
89
81
164
200
86
11
198
140
222
36
177
182
234
87
106
71
72
88
42
91
228
119
128
207
101
102
117
80
145

9
2
3
8
2
11
5
11
11
5
10
11
11
9
11
5
5
9
11
5
2
11
9
11
3
9
9
11
5
7
5
5
5
3
5
11
8
8
11
6
6
8
5
9

Photodegradation of Pesticides

131

Directory of pesticide chemical structures (Continued).


Pesticide
Chlorsulfuron
Chlorthiamid
Cinmethylin
CNP (Chlornitrofen)
Coumaphos
Cyanofenphos
Cyanophos
Cyfluthrin
Cyhalothrin
Cypermethrin
Cyphenothrin
Cyprodinil
2,4-D
DDD
DDE
DDOD
DDT
DTP
Deltamethrin
Desmedipham
Diafenthiuron
Diazinon
Dicamba
Dichlobenil
Dichlofluanid
Dichlorvos
Diclofop-methyl
Dicofol
Dicrotophos
Dieldrin
Diflubenzuron
Dimethoate
Dimetilam
Diniconazole-M
Dinobuton
Dinoseb
Dioxabenzofos
Diphenamid
Diquat
Disulfoton
Ditalimfos
Diuron
ETU
Edifenphos
Endosulfan

Identification number

Appendix number

96
49
249
217
147
180
137
20
21
19
14
210
1
132
131
113
130
221
22
83
57
144
118
116
205
153
5
133
157
123
59
165
75
190
203
202
183
50
226
163
178
53
95
173
125

6
3
11
11
9
9
9
2
2
2
2
11
1
8
8
7
8
11
2
5
4
9
8
8
11
9
1
8
9
8
4
9
5
10
11
11
9
3
11
9
9
4
5
9
8

132

T. Katagi
Directory of pesticide chemical structures (Continued).
Pesticide
Endrin
Esfenvalerate
Ethalfluralin
Ethiofencarb
Ethion
Ethirimol
Ethoxyquin
Etofenprox
Famoxadone
Fenarimol
Fenchlorphos
Fenitrothion
Fenobucarb
Fenothiocarb
Fenpropathrin
Fenpropimorph
Fenthion
Fentin acetate
Fenuron
Fenvalerate
Fipronil
Florasulam
Fluazifop-butyl
Fluchloralin
Flucythrinate
Fludioxinil
Flumetralin
Flumeturon
Fluotrimazole
Flutolanil
Fluvalinate
Folpet
Formothion
Guazatine
Haloxyfop
Heptachlor
Hexachlorobenzene
Hexaconazole
Imazapyr
Imazaquin
Imazethapyr
Imidacloprid
Iodofenphos
Iprodione
Isofenphos

Identification number

Appendix number

126
28
237
69
184
213
219
32
111
239
139
138
62
85
24
227
143
241
51
27
220
48
6
236
29
208
233
55
195
39
30
107
166
245
8
127
115
192
229
230
231
215
141
109
175

8
2
11
5
9
11
11
2
7
11
9
9
5
5
2
11
9
11
4
2
11
3
1
11
2
11
11
4
10
3
2
7
9
11
1
8
8
10
11
11
11
11
9
7
9

Photodegradation of Pesticides

133

Directory of pesticide chemical structures (Continued).


Pesticide
Isolan
Isoprothiolane
Isoproturon
Isoxaben
Isoxathion
Kadethrin
Lenacil
Leptophos
Lindane
Linuron
MAB1a
MBC
MCPA
Malathion
Maneb
Mecoprop
Mepronil
Metalaxyl
Methazole
Methidathion
Methiocarb
Methomyl
Methoprene
Methoxychlor
Metolachlor
Metolcarb
Metribuzin
Metsulfuron
Metsulfuron methyl
Mevinphos
Mexacarbate
Mobam
Molinate
Monocrotophos
Monuron
NAA
NMH
Naproanilide
Napropamide
Naptalam
Niclosamide
Nitrofen
Norflurazon
Oryzalin
Oxamyl

Identification number

Appendix number

77
204
56
46
151
16
199
181
114
54
251
93
3
167
94
4
38
37
112
171
68
70
246
134
35
61
201
97
98
155
67
73
90
156
52
247
242
41
47
44
40
216
214
235
74

5
11
4
3
9
2
11
9
8
4
11
5
1
9
5
1
3
3
7
9
5
5
11
8
3
5
11
6
6
9
5
5
5
9
4
11
11
3
3
3
3
11
11
11
5

134

T. Katagi
Directory of pesticide chemical structures (Continued).
Pesticide
Oxycarboxin
Oxyfluorfen
PBacid
Paraquat
Parathion
Parathion-methyl
Penconazole
Pendimethalin
Pentachlorophenol
Perfluidone
Permethrin
Phenmec
Phenmedipham
Phenothrin
Phenthoate
Phorate
Phosalone
Phosmet
Phosphamidon
Photo-dieldrin
Phoxim
Picloram
Pirimicarb
Potasan
Prochloraz
Procymidone
Profenofos
Prometone
Propachlor
Propaphos
Propham
Propiconazole
Propoxur
Propyzamide
Pyrazophos
Pyrethrin-I
Pyridaphenthion
Pyrimethanil
Pyrolan
Quinalophos
Resmethrin
Rimsulfuron
S-2571
S-Bioallethrin
Sethoxydim

Identification number

Appendix number

43
218
243
225
135
136
193
238
121
206
17
60
84
13
168
162
170
172
158
124
152
120
78
146
240
108
160
187
33
159
79
191
65
45
149
9
150
209
76
148
15
103
176
12
223

3
11
11
11
9
9
10
11
8
11
2
5
5
2
9
9
9
9
9
8
9
8
5
9
11
7
9
10
3
9
5
10
5
3
9
2
9
11
5
9
2
6
9
2
11

Photodegradation of Pesticides

135

Directory of pesticide chemical structures (Continued).


Pesticide
Simazine
Spinosyn A
Sulprofos
2,4,5-T
2,3,7,8-TCDD
Tefluthrin
Terbacil
Tetrachlovinphos
Tetramethrin
Thiabendazole
Thiadiazuron
Thifensulfuron
Thifensulfuron methyl
Thiophanate-methyl
Tolclofos-methyl
Tralocythrin
Tralomethrin
Triadimefon
Triadimenol
Triasulfuron
Tribenuron methyl
Trichlorfon
Triclopyr
Trifluralin
Trimethacarb
Triticonazole
Vinclozoline
Warfarine
Xylylcarb

Identification number

Appendix number

186
252
161
2
129
18
197
154
23
212
58
104
105
92
142
26
25
188
189
100
99
179
7
232
64
194
110
248
63

10
11
9
1
8
2
11
9
2
11
4
6
6
5
9
2
2
10
10
6
6
9
1
11
5
10
7
11
5

Pesticide
2,4-D
2,4,5-T
MCPA
Mecoprop
Diclofop-methyl
Fluazifop-butyl
Triclopyr
Haloxyfop

No.
1
2
3
4
5
6
7
8

C
C
C
C
C
C
2-N
C

X
2,4-Cl2
2,4,5-Cl3
4-Cl-2-CH3
4-Cl-2-CH3
4-(2,4-Cl2-Phenoxy)
5-CF3-2-Pyridyloxy
3,5,6-Cl3
3-Cl-5-CF3-2-Pyridyloxy

R1

Appendix 1. Chemical structures of aryloxyalkanoates.

H
H
H
CH3
CH3
CH3
H
CH3

R2
H
H
H
H
CH3
n-C4H9
H
H

R3

136
T. Katagi

Insecticide
Pyrethrin-I
Allethrin
Bioallethrin
S-Bioallethrin
Phenothrin
Cyphenothrin
Resmethrin
Kadethrin
Permethrin
Tefluthrin
Cypermethrin
Cyfluthrin
Cyhalothrin
Deltamethrin
Tetramethrin

No.
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23

A1
A1
A1
A1
A1
A1
A1
A2
A1
A1
A1
A1
A1
A1
A1

R1
CH3
CH3
CH3
CH3
CH3
CH3
CH3

Cl
Cl
Cl
Cl
Cl
Br
CH3

X
CH3
CH3
CH3
CH3
CH3
CH3
CH3

Cl
CF3
Cl
Cl
CF3
Br
CH3

Y
1R-trans
1RS-cis,trans
1R-trans
1R-trans
1RS-cis,trans
1RS-cis,trans
1RS-cis,trans
E-1R-cis
1RS-cis,trans
Z-1RS-cis
1RS-cis,trans
1RS-cis,trans
Z-1RS-cis
1R-cis
1RS-cis,trans

Configuration

Appendix 2. Chemical structures of pyrethroids.

B1
B1
B1
B1
B2
B2
B3
B3
B2
B2
B2
B2
B2
B2
B4

R2

H
CN

H
H
CN
CN
CN
CN

Z
Z-S
RS
RS
S

RS

RS
RS
RS
S

Configuration

CH=CH2
H
H
H
3-Phenoxy
3-Phenoxy

3-Phenoxy
2,3,5,6-Tetrafluoro-4-methyl
3-Phenoxy
4-Fluoro-3-phenoxy
3-Phenoxy
3-Phenoxy

R3

Photodegradation of Pesticides
137

24
25
26
(28)/27
29
30
31

No.
Fenpropathrin
Tralomethrin
Tralocythrin
(Es)fenvalerate
Flucythrinate
Fluvalinate
Acrinathrin

Insecticide

Appendix 2. (Continued).

A3
A4
A4
A5
A5
A5
A1

R1

Br
Cl

NH
H

4-Cl
4-OCHF2
2-Cl-4-CF3
COOCH(CF3)2

1R-cis
1R-cis
(2S) 2RS
2S
2R
Z-1R-cis

Configuration
B2
B2
B2
B2
B2
B2
B2

R2
CN
CN
CN
CN
CN
CN
CN

Z
RS
S
S
(S) RS
RS
RS
S

Configuration

3-Phenoxy
3-Phenoxy
3-Phenoxy
3-Phenoxy
3-Phenoxy
3-Phenoxy
3-Phenoxy

R3

138
T. Katagi

Pesticide
Propachlor
Alachlor
Metolachlor
Butachlor
Metalaxyl
Mepronil
Flutolanil
Niclosamide
Naproanilide
Carboxin
Oxycarboxin

No.
33
34
35
36
37
38
39
40
41
42
43

H
2,6-(C2H5)2
2-C2H5-6-CH3
2,6-(C2H5)2
2,6-(CH3)2
3-O-iso-C3H7
3-O-iso-C3H7
2-Cl-4-NO2
H
H
H

R1
CH2Cl
CH2Cl
CH2Cl
CH2Cl
CH2OCH3
2-CH3Ph
2-CF3Ph
3-Cl-6-OHPh
CH(CH3)O-2-Naphthyl
A1
A2

R2

Appendix 3. Chemical structures of anilides and amides.

CH(CH3)2
CH2OCH3
CH(CH3)CH2OCH3
CH2O(CH2)3CH3
CH(CH3)COOCH3
H
H
H
H
H
H

R3

Photodegradation of Pesticides
139

Pesticide
Fenuron
Monuron
Diuron
Linuron
Flumeturon
Isoproturon
Diafenthiuron

No.
51
52
53
54
55
56
57

O
O
O
O
O
O
S

X
H
4-Cl
3,4-Cl2
3,4-Cl2
3-CF3
4-iso-C3H7
2,6-(iso-C3H7)2-4-OPh

R1
CH3
CH3
CH3
OCH3
CH3
CH3
tert-C4H9

R2

Appendix 4. Chemical structures of ureas and benzoylureas.

CH3
CH3
CH3
CH3
CH3
CH3
H

R3

140
T. Katagi

Pesticide
Phenmec
Metolcarb
Fenobucarb
Xylylcarb
Trimethacarb
Propoxur
Aminocarb
Mexacarbate
Methiocarb
Ethiofencarb
Methomyl
Carbaryl
Carbofuran
Mobam
Oxamyl
Dimetilam
Pyrolan
Isolan
Pirimicarb
Propham
Chlorpropham
Benomyl
Asulam
Desmedipham
Phenmedipham

No.
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84

CH3
CH3
CH3
CH3
CH3
CH3
CH3
CH3
CH3
CH3
CH3
CH3
CH3
CH3
CH3
CH3
CH3
CH3
CH3
Ph
3-ClPh
A7
A8
Ph
3-CH3Ph

R1

Appendix 5. Chemical structures of carbamates.

H
H
H
H
H
H
H
H
H
H
H
H
H
H
H
CH3
CH3
CH3
CH3
H
H
H
H
H
H

R2
Ph
3-CH3Ph
2-sec-C4H9Ph
3,4-(CH3)2Ph
3,4,5 (or 2,3,5)-(CH3)3Ph
2-O-iso-C3H7Ph
3-CH3-4-N(CH3)2Ph
3,5-(CH3)2-4-N(CH3)2Ph
3,5-(CH3)2-4-SCH3Ph
2-CH2SC2H5-Ph
N=C(CH3)SCH3
Naphth-1-yl
A1
A2
N=C(SCH3)C(O)N(CH3)2
A3
A4
A5
A6
iso-C3H7
iso-C3H7
CH3
CH3
3-NHCO2C2H5Ph
3-NHCO2CH3Ph

R3
O
O
O
O
O
O
O
O
O
O
O
O
O
O
O
O
O
O
O
O
O
O
O
O
O

X
O
O
O
O
O
O
O
O
O
O
O
O
O
O
O
O
O
O
O
O
O
O
O
O
O

Photodegradation of Pesticides
141

Pesticide
Fenothiocarb
Benthiocarb
Butylate

No.
85
86
87

Appendix 5. (Continued).

CH3
C2H5
(CH3)2CHCH2

R1
CH3
C2H5
(CH3)2CHCH2

R2
(CH2)4OPh
CH2-4-ClPh
C2H5

R3
O
O
O

S
S
S

142
T. Katagi

Appendix 5. (Continued).

Photodegradation of Pesticides

143

Herbicide
Chlorsulfuron
Metsulfuron
Metsulfuron methyl
Tribenuron methyl
Triasulfuron
Chlorimuron
Chlorimuron ethyl
Rimsulfuron
Thifensulfuron
Thifensulfuron methyl

No.
96
97
98
99
100
101
102
103
104
105

C
C
C
C
C
C
C
C
A1
A2

X
C
C
C
C
C
C
C
N

R1

R2
Cl
COOH
COOCH3
COOCH3
OCH2CH2Cl
COOH
COOC2H5
SO2C2H5

Aryl

Appendix 6. Chemical structures of sulfonylureas.

N
N
N
N
N
C
C
C
N
N

Z
H
H
H
CH3
H
H
H
H
H
H

R3
OCH3
OCH3
OCH3
OCH3
OCH3
Cl
Cl
OCH3
OCH3
OCH3

R4
CH3
CH3
CH3
CH3
CH3
OCH3
OCH3
OCH3
CH3
CH3

R5

144
T. Katagi

Appendix 7. Chemical structures of cyclic dicarboximides and related pesticides.

Photodegradation of Pesticides
145

Appendix 8. Chemical structures of organochlorine pesticides.

146
T. Katagi

Appendix 8. (Continued).

Photodegradation of Pesticides

147

Pesticide
Parathion
Parathion-methyl
Cyanophos
Fenitrothion
Fenchlorphos
Bromophos
Iodofenphos
Tolclofos-methyl
Fenthion
Diazinon
Chlorpyrifos
Potasan
Coumaphos
Quinalophos
Pyrazophos
Pyridaphenthion
Isoxathion
Phoxim
Dichlorvos
Tetrachlovinphos
Mevinphos
Monocrotophos
Dicrotophos
Phosphamidon
Propaphos

No.
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159

OC2H5
OCH3
OCH3
OCH3
OCH3
OCH3
OCH3
OCH3
OCH3
OC2H5
OC2H5
OC2H5
OC2H5
OC2H5
OC2H5
OC2H5
OC2H5
OC2H5
OCH3
OCH3
OCH3
OCH3
OCH3
OCH3
O-nC3H7

R1
OC2H5
OCH3
OCH3
OCH3
OCH3
OCH3
OCH3
OCH3
OCH3
OC2H5
OC2H5
OC2H5
OC2H5
OC2H5
OC2H5
OC2H5
OC2H5
OC2H5
OCH3
OCH3
OCH3
OCH3
OCH3
OCH3
O-nC3H7

R2

R3
O(4-NO2Ph)
O(4-NO2Ph)
O(4-CNPh)
O(3-CH3-4-NO2Ph)
O(2,4,5-Cl3Ph)
O(2,5-Cl2-4-BrPh)
O(2,5-Cl2-4-IPh)
O(2,6-Cl2-4-CH3Ph)
O(3-CH3-4-SCH3Ph)
A1
O(3,5,6-Cl3-pyridin-2-yl)
A2
A3
A4
A5
A6
A7
ON=C(CN)Ph
OCH=CCl2
OC(=CHCl)(2,4,5-Cl3Ph)
OC(CH3)=CHCOOCH3
OC(CH3)=CHC(O)NHCH3
OC(CH3)=CHC(O)N(CH3)2
OC(CH3)=CClC(O)N(C2H5)2
O(4-SCH3Ph)

Appendix 9. Chemical structures of organophosphorus pesticides.

S
S
S
S
S
S
S
S
S
S
S
S
S
S
S
S
S
S
O
O
O
O
O
O
O

148
T. Katagi

Pesticide
Profenofos
Sulprofos
Phorate
Disulfoton
Bensulide
Dimethoate
Formothion
Malathion
Phenthoate
Azinphos-methyl
Phosalone
Methidathion
Phosmet
Edifenphos
Acephate
Isofenphos
S-2571
Butamifos
Ditalimfos
Trichlorfon
Cyanofenphos
Leptophos
Butonate

No.
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182

Appendix 9. (Continued).

OC2H5
OC2H5
OC2H5
OC2H5
O-iso-C3H7
OCH3
OCH3
OCH3
OCH3
OCH3
OC2H5
OCH3
OCH3
OC2H5
OCH3
OC2H5
OC2H5
OC2H5
OC2H5
OCH3
OC2H5
OCH3
OCH3

R1
S-nC3H7
S-nC3H7
OC2H5
OC2H5
O-iso-C3H7
OCH3
OCH3
OCH3
OCH3
OCH3
OC2H5
OCH3
OCH3
SPh
NHC(O)CH3
NH-iso-C3H7
NH-iso-C3H7
NH-sec-C4H9
OC2H5
OCH3
O(4-CNPh)
O(2,5-Cl2-4-BrPh)
OCH3

R2
O(4-Br-2-ClPh)
O(4-SCH3)Ph
SCH2SC2H5
SCH2CH2SC2H5
SCH2CH2NHS(O)2Ph
SCH2C(O)NHCH3
SCH2C(O)N(CHO)CH3
SCH(COOC2H5)CH2COOC2H5
SCH(COOC2H5)Ph
A8
A9
A10
A11
SPh
SCH3
O(2-C(O)O-iso-C3H7Ph)
O(3-CH3-6-NO2Ph)
O(3-CH3-6-NO2Ph)
A12
CH(OH)CCl3
Ph
Ph
CH(CCl3)OC(=O)C3H7

R3
O
S
S
S
S
S
S
S
S
S
S
S
S
O
O
S
S
S
S
O
S
S
O

Photodegradation of Pesticides
149

Appendix 9. (Continued).

150
T. Katagi

Pesticide
Atrazine
Simazine
Prometone

No.
185
186
187

Cl
Cl
OCH3

R1
CH(CH3)2
CH2CH3
CH(CH3)2

R2
CH2CH3
CH2CH3
CH(CH3)2

R3

Appendix 10. Chemical structures of triazines and azoles.

Photodegradation of Pesticides
151

Appendix 11. Chemical structures of miscellaneous pesticides.

152
T. Katagi

Appendix 11. (Continued).

Photodegradation of Pesticides

153

Appendix 11. (Continued).

154
T. Katagi

Appendix 11. (Continued).

Photodegradation of Pesticides

155

Appendix 11. (Continued).

156
T. Katagi

Photodegradation of Pesticides

157

References
Aaron JJ, Kaleel EM, Winefordner JD (1979) Comparative study of low-temperature and
room-temperature phosphorescence characteristics of several pesticides. J Agric Food
Chem 27:12331237.
Abbott CK, Sorensen DL, Sims RC (1992) Use and efficiency of ethylene glycol monomethyl ether and monoethanolamine to trap volatilized [7-14C]naphthalene and 14CO2.
Environ Toxicol Chem 11:181185.
Abdel-Malik MM, de Mayo P (1984) Surface photochemistry: the amide photo-Fries
rearrangement. Can J Chem 62:12751278.
Abdel-Wahab AM, Casida JE (1967) Photooxidation of two 4-dimethylaminoaryl methylcarbamate insecticides (zectran and matacil) on bean foliage and of alkylaminophenyl methylcarbamates on silica gel chromatoplates. J Agric Food Chem 15:479487.
Abdel-Wahab AM, Kuhr RJ, Casida JE (1966) Fate of 14C-carbonyl-labeled aryl methylcarbamate insecticide chemicals in and on bean plants. J Agric Food Chem 14:290
298.
Abdou WM, Mahran MR, Sidky MM, Wamhoff H (1988) Photochemistry of pesticides.
13. Some photoreactions of O,O-diethyl O-(4-methyl-2-oxo-2H-1-benzopyran-7-yl)
phosphorothioate (Potasan). Phosphorus Sulfur Relat Elem 39:199203.
Abe Y, Tsuda K, Fujita Y (1972) Studies on pyrethroidal compounds. Part III. Photostability of pyrethroidal compounds. Botyu-Kagaku 37:102111.
Adams JD, Iwata Y, Gunther FA (1977) Worker environment research. V. Effect of soil
dusts on dissipation of paraoxon dislodgeable residues on citrus foliage. Bull Environ
Contam Toxicol 18:445451.
Addison JB, Silk PJ, Unger I (1974) The photochemical reactions of carbamates. II. The
solution photochemistry of matacil (4-dimethylamino-m-tolyl-N-methylcarbamate)
and landrin (3,4,5-trimethylphenyl-N-methylcarbamate). Bull Environ Contam Toxicol 11:250255.
Addison JB, Semeluk GP, Unger I (1977) The luminescence properties of pesticides.
I. Fluorescencing and phosphorescing carbamates. J Lumin 15:323339.
Aguer JP, Richard C (1996a) Reactive species produced on irrdiation at 365 nm of
aqueous solutions of humic acids. J Photochem Photobiol A Chem 93:193198.
Aguer JP, Richard C (1996b) Transformation of fenuron induced by photochemical excittion of humic acids. Pestic Sci 46:151155.
Aguer JP, Richard C (1999) Influence of the excitation wavelength on the photoinductive
properties of humic substances. Chemosphere 38:22932301.
Aguer JP, Richard C, Andreux F (1997) Comparison of the photoinductive properties of
commercial, synthetic and soil-extracted humic substances. J Photochem Photobiol A
Chem 103:163168.
Aguer JP, Blachere F, Boule P, Garaudee S, Guillard C (2000) Photolysis of dicamba
(3,6-dichloro-2-methoxybenzoic acid) in aqueous solution and dispersed on solid supports. Int J Photoenergy 2:8186.
Aguer JP, Trubetskaya O, Trubetskoj O, Richard C (2001) Photoinductive properties of
soil humic acids and their fractions obtained by tandem size exclusion chromatographypolyacrylamide gel electrophoresis. Chemosphere 44:205209.
Aguer JP, Richard C, Trubetskaya O, Trubetskoj O, Leve`gue J, Andreux F (2002) Photoinductive efficiency of soil extracted humic and fulvic acids. Chemosphere 49:259
262.

158

T. Katagi

Albanis TA, Bochicchio D, Bufo SA, Cospito I, DAuria M, Lekka M, Scrano L (2002)
Surface adsorption and photo-reactivity of sulfonylurea herbicides. Int J Anal Chem
82:561569.
lenius CM, Vogelmann TC, Bornman JF (1995) A three-dimensional representation of
A
the relationship between penetration of UV-B radiation and UV-screening pigments
in leaves of Brassica napus. New Phytol 131:297302.
Allebone JE, Hamilton RJ, Bryce TA, Kelly W (1971) Anthraquinone in plant surface
waxes. Experientia (Basel) 27:1314.
Allmaier GM, Schmid ER (1985) Effects of light on the organophosphorus pesticides
bromophos and iodofenphos and their main degradation products examined in rain
water and on soil surface in a long-term study. J Agric Food Chem 33:9092.
Allmaier GM, Fogy I, Heinisch G, Schmid ER (1984) Photolysis of three organophosphorus pesticides in rain water and soil surface. In: Frigerio A, Milon H (eds) Chromatography and Mass Spectrometry in Nutrition Science and Food Safety. Elsevier,
Amsterdam, pp 115125.
Ando M, Iwasaki Y, Nakagawa M (1975) Metabolism of isoxathion, O,O-dimethyl O(5-phenyl-3-isoxazolyl) phosphorothioate in plants. Agric Biol Chem 39:21372143.
Archer TE, Stokes JD, Bringhurst RS (1977) Fate of carbofuran and its metabolites on
strawberries in the environment. J Agric Food Chem 25:536541.
Argauer RJ (1980) Fluorescence and ultraviolet absorbance of pesticides and naturally
occurring chemicals in agricultural products after HPLC separation on a bonded-CN
polar phase. In: Harvey J Jr, Zweig G (eds) Pesticide Analytical Methodology. ACS
Symposium Series 136. American Chemical Society, Washington, DC, pp 103126.
Avato P, Bianchi G, Pogna N (1990) Chemosystematics of surface lipids from maize
and some related species. Phytochemistry (Oxf) 29:15711576.
Avnir D, Johnston LJ, DeMayo P, Wong SK (1981) Surface photochemistry: radical pair
combination on a silica gel surface and in micelles. J Chem Soc Chem Commun
958959.
Awad TM, Vinson SB, Brazzel JR (1967) Effect of environmental and biological factors
on persistence of malathion applied as ultra-low-volume or emulsifiable concentrate
to cotton plants. J Agric Food Chem 15:10091013.
Bailey GW, Karickhoff SW (1973) UV-VIS spectroscopy in the characterization of clay
mineral surfaces. Anal Lett 6:4349.
Baker EA (1982) Chemistry and morphology of plant epicuticular waxes. In: Cutler DF,
Alvin KL, Price CE (eds) The Plant Cuticles. Linnean Society Symposium Series 10
(Plant Cuticle). Academic Press, New York, pp 139165.
Baker EA, Hunt G (1981) Developmental changes in leaf epicuticular waxes in relation
to foliar penetration. New Phytol 88:731747.
Baker EA, Bukovac MJ, Flore JA (1979) Ontogenic variations in the composition of
peach leaf wax. Phytochemistry (Oxf) 18:781784.
Baker EA, Bukovac MJ, Hunt GM (1982) Compositions of tomato fruit cuticle as related
to fruit growth and development. In: Cutler DF, Alvin KL, Price CE (eds) The Plant
Cuticles. Linnean Society Symposium Series 10 (Plant Cuticle). Academic Press,
New York, pp 3344.
Baker EA, Hunt GM, Stevens PJG (1983) Studies of plant cuticle and spray droplet
interactions: a fresh approach. Pestic Sci 14:645658.
Baker EA, Procopiou J, Hunt GM (1998) The cuticles of citrus species. Composition of
leaf and fruit waxes. J Sci Food Agric 26:10931101.

Photodegradation of Pesticides

159

Balkaya N (2003) Photocatalytic degradation technology for pesticide elimination from


aquatic media available. Energy Educ Sci Technol 10:7380.
Balmer ME, Sulzberger B (1999) Atrazine degradation in irradiated iron/oxalate systems:
effects of pH and oxalate. Environ Sci Technol 33:24182424.
Balmer ME, Goss KU, Schwarzenbach (2000) Photolytic transformation of organic pollutants on soil surfaces: an experimental approach. Environ Sci Technol 34:1240
1245.
Bandal SK, Casida JE (1972) Metabolism and photoalteration of 2-sec-butyl-4,6-dinitrophenol (DNBP herbicide) and its isopropyl carbonate derivative (dinobuton acaricide). J Agric Food Chem 20:12351245.
Banerjee K, Dureja P (1999) Phototransformation of quinalphos on clay surfaces. Toxicol Environ Chem 68:475480.
Barnes RD, Bull AT, Poller RC (1973) Studies on persistence of the organotin fungicide
fentin acetate (triphenyltin acetate) in the soil and on surfaces exposed to light. Pestic
Sci 4:305317.
Barraclough D, Nye PH (1979) The effect of molecular size on diffusion characteristics
in soil. J Soil Sci 30:2942.
Barta IC, Komives T (1984) Gas-liquid chromatographic method for the rapid analysis
of the epicuticular wax composition of plants. J Chromatogr 287:438441.
Basham GW, Lavy TL (1987) Microbial and photolytic dissipation of imazaquin in soil.
Weed Sci 35:865870.
Baude FJ, Gardiner JA, Han JCY (1973) Characterization of residues on plants following
foliar spray applications of benomyl. J Agric Food Chem 21:10841090.
Bauer H, Schonherr J (1992) Determination of mobilities of organic compounds in plant
cuticles and correlation with molar volumes. Pestic Sci 35:111.
Beigel C, Barriaso E, Di Pietro L (1997) Time dependency of triticonazole fungicide
sorption and consequences for diffusion in soil. J Environ Qual 26:15031510.
Belding RD, Blankenship SM, Young E, Leidy RB (1998) Composition and variability
of epicuticular waxes in apple cultivars. J Am Soc Hortic Sci 123:348356.
Bengston C, Larsson S, Liljenberg C (1978) Effects of water stress on cuticular transpiration rate and amount and composition of epicuticular wax in seedlings of six oat
varieties. Physiol Plant 44:319324.
Benson WR (1971) Photolysis of solid and dissolved dieldrin. J Agric Food Chem 19:
6672.
Benson WR, Lombardo P, Egry IJ, Ross IJ Jr, Barron RP, Mastbrook DW, Hansen EA
(1971) Chlordane photoalteration products: their preparations and identification.
J Agric Food Chem 19:857862.
Bentson KP (1990) Fate of xenobiotics in foliar pesticide deposits. Rev Environ Contam
Toxicol 114:125161.
Berenbaum MR, Larson RA (1988) Flux of singlet oxygen from leaves of phototoxic
plants. Experientia (Basel) 44:10301032.
Bertino DJ, Zepp RG (1991) Effects of solar radiation on manganese oxide reactions
with selected organic compounds. Environ Sci Technol 25:12671273.
Beynon KI, Wright AN (1969) Breakdown of the insecticide Gardona on plants and in
soils. J Sci Food Agric 20:250256.
Beynon KI, Wright AN (1972) The breakdown of [14C] monocrotophos insecticide on
maize, cabbage and apple. Pestic Sci 3:277292.
Bhattacharjee AK, Dureja P (1999) Light induced transformation of tribenuron-methyl.
Chemosphere 38:741749.

160

T. Katagi

Bhattacharjee AK, Dureja P (2002) Light-induced transformation of tribenuron-methyl


on glass, soil and plant surfaces. J Environ Sci Health B37:131140.
Bianchi G (1995) Plant waxes. In: Hamilton RJ (ed) Waxes: Chemistry, Molecular Biology and Functions, vol 6. The Oily Press, Dundee, pp 175222.
Bobe A, Cooper JF, Coste CM, Muller MA (1998a) Behavior of fipronil in soil under
Sahelian plain field conditions. Pestic Sci 52:275281.
Bobe A, Meallier P, Cooper JF, Coste CM (1998b) Kinetics and mechanisms of abiotic
degradation of fipronil (hydrolysis and photolysis). J Agric Food Chem 46:2834
2839.
Bornman JF, Vogelman TC (1991) Effect of UV-B radiation on leaf optical properties
measured with fiber optics. J Exp Bot 42:547554.
Bowman MC, Beroza M (1967a) Spectra and analysis of insecticide synergists and related compounds containing the methylenedioxyphenyl group by spectrophotofluorometry (SPF) and spectrophotophosphorimetry (SPF). Residue Rev 17:122.
Bowman MC, Beroza M (1967b) Spectrophotofluorescent and spectrophotophosphorescent data on insecticidal carbamates and the analysis of five carbamates in milk by
spectrophotofluorometry. Residue Rev 17:2334.
Bowman MC, Leuck DB (1971) Determination and persistence of phoxim and its oxygen
analog in forage corn and grass. J Agric Food Chem 19:12151218.
Breithaupt DE, Schwack W (2000) Photoinduced addition of the fungicide anilazine to
cyclohexene and methyl oleate as model compounds of plant cuticle constituents.
Chemosphere 41:14011406.
Breugem P, van Noort P, Velberg S, Wondergem E, Zijlstra J (1986) Steady-state concentrations of the phototransient hydrated electron in natural waters. Chemosphere
15:717724.
Brown HM, Brattsten LB, Lilly DE, Hanna PJ (1993) Metabolic pathways and residue
levels of thifensulfuron methyl in soybeans. J Agric Food Chem 41:17241730.
Buchenauer H (1975) Differences in light stability of some carboxylic acid anilide fungicides in relation to their applicability for seed and foliar treatment. Pestic Sci 6:
525535.
Buchenauer H, Edington LV, Grossmann F (1973) Photochemical transformation of thiophanate-methyl and thiophanate to alkyl benzimidazol-2-yl carbamates. Pestic Sci 4:
343348.
Bukovac MJ, Flore JA, Baker EA (1979) Peach leaf surfaces: changes in wettability,
retention, cuticular permeability, and epicuticular wax chemistry during expansion
with special reference to spray application. J Am Soc Hortic Sci 104:611617.
Bull DL (1979) Fate and efficacy of acephate after application to plants and insects.
J Agric Food Chem 27:268272.
Bull DL, Ivie GW (1976) Fate of diflubenzuron in cotton, soil and rotational crops.
J Agric Food Chem 26:515520.
Bull DL, Lindquist DA (1964) Metabolism of 3-hydroxy-N,N-dimethylcrotonamide dimethyl phosphate by cotton plants, insects and rats. J Agric Food Chem 12:310317.
Bull DL, Lindquist DA, Grabbe RR (1967) Comparative fate of the geometric isomers
of phosphamidon in plants and animals. J Econ Entomol 60:332241.
Bullivant MJ, Pattenden G (1973) Photolysis of bio-allethrin. Tetrahedron Lett 3679
3680.
Bullivant MJ, Pattenden G (1976) Photochemistry of 2-(prop-2-enyl)cyclopent-2-enones.
J Chem Soc Perkin Trans I:249256.

Photodegradation of Pesticides

161

Burghardt M, Schreiber L, Riederer M (1998) Enhancement of the diffusion of active


ingredients in barley leaf cuticular wax by monodisperse alcohol ethoxylates. J Agric
Food Chem 46:15931602.
Burkhard N, Guth JA (1979) Photolysis of organophosphorus insecticides on soil surfaces. Pestic Sci 10:313319.
Burrows HD, Canle M, Santaballa JA, Steenken S (2002) Reaction pathways and mechanisms of photodegradation of pesticides. J Photochem Photobiol B Biol 67:71108
Busch MA, Franklin TC (1979) Nature of the copper(I) carbonyl formed by acid cuprous
chloride in the presence of carbon monoxide. Inorg Chem 18:521524.
Caboni P, Cabras M, Angioni A, Russo M, Cabras P (2002) Persistence of azadirachtin
residues on olives after field treatment. J Agric Food Chem 50:34913494.
Cabras P, Spanedda L, Cabitza F, Cubeddu M, Martini MG, Brandolini V (1990) Pirimicarb and its metabolite residues in lettuce. Influence of cultural environment. J Agric
Food Chem 38:879882.
Cabras P, Angioni A, Garau VL, Melis M, Pirisi FM, Minelli EV, Cabitza F, Cubeddu
M (1997a) Fate of some new fungicides (cyprodinil, fludioxinil, pyrimethanil and
tebuconazole) from vine to wine. J Agric Food Chem 45:27082710.
Cabras P, Angioni A, Garau VL, Melis M, Pirisi FM, Minelli EV (1997b) Effect of
epicuticular waxes of fruits on the photodegradation of fenthion. J Agric Food Chem
45:36813683.
Cabras P, Angioni A, Garau VL, Melis M, Pirisi FM, Espinoza J, Mendoza A, Cabitza
F, Pala M, Brandolini V (1998) Fate of azoxystrobin, fluazinam, kresoxim-methyl,
mepanipyrim, and tetraconazole from vine to wine. J Agric Food Chem 46:3249
3251.
Caine M, Dyer G, Holder JV, Osborne BN, Matear WA, McCabe RW, Mobbs D, Richardson S, Wang L (1999) The use of clays as sorbents and catalysts. In: Misaelides
P, Macasek F, Pinnavaia TJ, Colella C (eds) Natural Microporous Materials in Environmental Technology. Kluwer, Dordrecht, pp 4969.
Calumpang SMF, Tejada AW, Magallona ED (1984) Photodecomposition on silica gel
plates of nine selected organophosphorus insecticide formulations. Philipp Entomol
6:191204.
Campbell JR, Penner D (1985a) Abiotic transformations of sethoxydim. Weed Sci 33:
435439.
Campbell JR, Penner D (1985b) Sethoxydim metabolism in monocotyledonous and dicotyledonus plants. Weed Sci 33:771773.
Canonica S, Jans U, Stemmler K, Hoigne (1995) Transformation kinetics of phenols in
water: photosensitization by dissolved natural organic material and aromatic ketones.
Environ Sci Technol 29:18221831.
Cape JN (1997) Photochemical oxidantswhat else is in the atmosphere besides ozone?
Phyton (Austria) 37:4558.
Carmichael I, Hug GL (1989) Spectroscopy and intramolecular photophysics of triplet
states. In: Scaiano JCD (ed) CRC Handbook of Organic Photochemistry, vol 1. CRC
Press, Boca Raton, FL, pp 369403.
Carruthers W, Johnstone RAW (1959) Composition of a paraffin wax fraction from
tobacco leaf and tobacco smoke. Nature (Lond) 184:11311132.
Casida JE, Gatterdam PE, Getzin Jr LW, Chapman RK (1956) Residual properties of the
systemic insecticide O,O-dimethyl 1-carbomethoxy-1-propen-2-yl phosphate. J Agric
Food Chem 4:236243.

162

T. Katagi

Cen YP, Bornman JF (1993) The effect of exposure to enhanced UV-B radiation on the
penetration of monochromatic and polychromatic UV-B radiation in leaves of Brassica napus. Physiol Plant 87:249255.
Cessna AJ, Muir DCG (1991) Photochemical transformations. In: Grover R, Cessna AJ
(eds) Environmental Chemistry of Herbicides, vol 2. CRC Press, Boca Raton, FL, pp
199263.
Chambers RW, Kearns DR (1969) Triplet states of some common photosensitizing dyes.
Photochem Photobiol 10:215219.
Chameides WL (1989) The chemistry of ozone deposition to plant leaves: Role of
ascorbic acid. Environ Sci Technol 23:595600.
Chattopadhyaya S, Dureja P (1991) Photolysis of flucythrinate. Pestic Sci 31:163173.
Chen J, Quan X, Yang F, Peijnenburg WJGM (2001) Quantitative structure-property
relationships on photodegradation of PCDD/Fs in cuticular waxes of laurel cherry
(Prunus laurocerasus). Sci Total Environ 269:163170.
Chen PH, Watts RJ (2000) Determination of rates of hydroxyl radical generation in
mineral-catalyzed Fenton-like oxidation. J Chin Inst Environ Eng 10:201208.
Chen YL, Casida JE (1969) Photodecomposition of pyrethrin I, allethrin, phthalthrin,
and dimethrin. J Agric Food Chem 17:208215.
Chen YL, Chen CC (1978) Photodecomposition of a herbicide, butachlor. J Pestic Sci 3:
143148.
Chen ZM, Zabik MJ, Leavitt RA (1984) Comparative study of thin film photodegradative
rates for 36 pesticides. Indian Eng Chem Prod Res Dev 23:511.
Cheng HM, Hwang DF (1996) Photodegradation of benthiocarb. Chem Ecol 12:91101.
Chesters G, Simsiman GV, Levy J, Alhajjar BJ, Fathulla RN, Harkin JM (1989) Environmental fate of alachlor and metolachlor. Rev Environ Contam Toxicol 110:174.
Chiba M, Kato S, Yamamoto I (1976) Metabolism of cyanox and surecide in bean
plants and degradation in soil. J Pestic Sci 1:179191.
Chou SS, Taniguchi E, Eto M (1980) Photodegradation of dialkyl 1,3-dithiolan-2-ylidenemalonates and some other pesticides on solid particles. Agric Biol Chem 44:1169
2177.
Choudhry GG (1981) Humic substances. Part II: Photophysical, photochemical and free
radical characterization. Toxicol Environ Chem 4:261295.
Choudhry GG (1984a) Photophysical and photochemical properties of soil and aquatic
humic materials. Residue Rev 92:59112.
Choudhry GG (1984b) Humic substances. Structural aspects, and photophysical, photochemical and free radical characteristics. In: Hutzinger O (ed) The Handbook of Environmental Chemistry, vol 1, part C. Springer-Verlag, Berlin, pp 124.
Choudhry GG, Webster GRB (1985) Protocol guidelines for the investigations of photochemical fate of pesticides in water, air, and soils. Residue Rev 96:80136.
Choudhury PP, Dureja P (1997a) Studies on photodegradation of chlorimuron-ethyl in
soil. Pestic Sci 51:201205.
Choudhury PP, Dureja P (1997b) Photolysis of chlorimuron-methyl. Toxicol Environ
Chem 63:7181.
Choudhury PP, Dureja P (1997c) Photolysis of chlorimuron-ethyl in benzene. Toxicol
Environ Chem 61:187193.
Chukwudebe A, March RB, Othman M, Fukuto TR (1989) Formation of trialkyl phosphorothioate esters from organophosphorus insecticides after exposure to either ultraviolet light or sunlight. J Agric Food Chem 37:539545.

Photodegradation of Pesticides

163

Cieslik S, Labatut A (1997) Ozone deposition on various surface types. Transp Chem
Transform Pollut Troposphere 4:225243.
Clark T, Watkins DAM (1986) Photolysis of triadimenol. Chemosphere 15:765770.
Clark T, Clifford DR, Deas AHB, Gendle P, Watkins DAM (1978) Photolysis, metabolism and other factors influencing the performance of triadimefon as a powdery mildew fungicide. Pestic Sci 9:497506.
Clark T, Watkins DAM, Weerasinghe DK (1983) Photolysis of fluotrimazole. Pestic Sci
14:449452.
Class TJ, Casida JE, Ruzo LO (1989) Photochemistry of etofenprox and three related
pyrethroids with ether, alkane, and alkene central linkage. J Agric Food Chem 37:
216222.
Clay VE, Fukuto TR (1984) Metabolism of carbosulfan in Valencia orange tree leaves
and fruits. Arch Environ Contam Toxicol 13:5362.
Clements P, Wells CHJ (1992) Soil sensitized generation of singlet oxygen in the photodegradation of bioresmethrin. Pestic Sci 34:163166.
Coats JR, Metcalf RL, Kapoor IP, Chio LC, Boyle PA (1979) Physical-chemical and
biological degradation studies on DDT analogues with altered aliphatic moieties.
J Agric Food Chem 27:10161022.
Cockell CS, Knowland J (1999) Ultraviolet radiation screening compounds. Biol Rev
74:311345.
Cole LM, Casida JE, Ruzo LO (1982) Comparative degradation of the pyrethroids tralomethrin, tralocythrin, deltamethrin, and cypermethrin on cotton and bean foliage.
J Agric Food Chem 30:916920.
Conceicao M, Mateus DA, da Silva AM, Burrows HD (1997) UV-visible absorption
spectra and luminescence of the pesticide fenarimol. Spectrochim Acta Part A Mol
Biomol Spectrosc 53:26792684.
Cookson RC, Dandegaonker SH (1955) Absorption spectra of ketones. Part III. The
long-wavelength band of -unsaturated ketones. J Chem Soc 16511654.
Cooper WJ (1989) Sunlight-induced photochemistry of humic substances in natural waters: major reactive species. In: Suffet IH, MacCarthy PM (eds) Aquatic Humic SubstancesInfluence on Fate and Treatment of Pollutants. Advances in Chemistry Series 219. American Chemical Society, Washington, DC, pp 333362.
Crosby DG (1983) Atmospheric reactions of pesticides. In: Miyamoto J, Kearney PC
(eds) Human Welfare and the Environment. Proceedings, 5th International Congress
on Pesticide Chemistry, vol 3. Pergamon Press, Oxford, pp 327332.
Crosby DG, Bowers JB (1985) Composition and photochemical reactions of a dimethylamine salt formulation of (4-chloro-2-methylphenoxy)acetic acid (MCPA). J Agric
Food Chem 33:569573.
Crosby DG, Wong AS (1976) Environmental degradation of 2,3,7,8-tetrachlorodibenzop-dioxin (TCDD). Science 195:13371338.
Crosley DR (1995) The measurement of OH and HO2 in the atmosphere. J Atmos Sci
52:32993314.
Crouch LS, Feely WF, Arison BH, VandenHeuvel WJA, Colwell LF, Stearns RA, Kline
WF, Wislocki PG (1991) Photodegradation of avermectin B1a thin films on glass.
J Agric Food Chem 39:13101319.
Crutzen PJ (1982) The global distribution of hydroxyl. In: Goldberg ED (ed) Atmospheric Chemistry. Springer-Verlag, Berlin, pp 313328.
Curran WS, Loux MM, Liebl RA, Simmons FW (1992) Photolysis of imidazolinone
herbicides in aqueous solutions and on soil. Weed Sci 40:143148.

164

T. Katagi

Da Silva JP, Da Silva AM, Khmelinskii IV, Martinho JMG, Vieira Ferreira LF (2001)
Photophysics and photochemistry of azole fungicides: triadimefon and triadomenol.
J Photochem Photobiol A Chem 142:3137.
Datta S, Walia S (1997) Photodegradation of buprofezin. Toxicol Environ Chem 60:
111.
Dauterman WC, Viado GB, Casida JE, OBrien RD (1960) Persistence of dimethoate
and metabolites following foliar application to plants. J Agric Food Chem 8:115119.
Day TA, Howells BW, Rice WJ (1994) Ultraviolet absorption and epidermal-transmittance spectra in foliage. Physiol Plant 92:207218.
Derek W, Georgi W, Grahl R (1979) Comparative degradation and metabolism of 32Plabeled butonate, trichlorfone and dichlorvos in crop plants. Biochem Physiol Pflanz
174:707722.
Dixon SR, Wells CHJ (1983) Dye-sensitized photo-oxidation of 2-dimethylamino-5,6dimethylpyrimidin-4-ol in aqueous solution. Pestic Sci 14:444448.
Dixon SR, Wells CHJ (1987) Chlorophyll sensitized photodegradation of 2-dimethylamino-5,6-dimethylpyrimidin-4-ol. Pestic Sci 21:155163.
Dodge AD, Knox JP (1986) Photosensitizers from plants. Pestic Sci 17:579586.
Donaldson SG, Miller GC (1996) Coupled transport and photodegradation of napropamide in soils undergoing evaporation from a shallow water table. Environ Sci Technol
30:924930.
Dorough HW, Witacre DM, Cardona RA (1973) Metabolism of the herbicide methazole
in cotton and beans, and fate of certain of its polar metabolites in rats. J Agric Food
Chem 21:797803.
Dorn HP, Brandenburger U, Brauers T, Hausmann M (1995) A new in situ laser longpath absorption instrument for the measurement of tropospheric OH radicals. J Atmos
Sci 52:33733380.
Drabek J, Boger M, Ehrenfreund J, Stamm E, Steinemann A, Alder A, Burckhardt U
(1992) New thioureas as insecticides. In: Crombie L (ed) Recent Advances in the
Chemistry of Insect Control, vol II. Special publication 79. Royal Society of Chemistry, Cambridge, pp 170183.
Draper WM, Casida JE (1985) Nitroxide radical adducts of nitrodiphenyl ether herbicides
and other nitroaryl pesticides with unsaturated cellular lipids. J Agric Food Chem 33:
103108.
Draper WM, Crosby DG (1981) Hydrogen peroxide and hydroxyl radical: intermediates
in indirect photolysis reactions in water. J Agric Food Chem 29:699702.
Draper WM, Crosby DG (1983) The photochemical generation of hydrogen peroxide in
natural waters. Arch Environ Contam Toxicol 12:121126.
Draper WM, Crosby DG (1984) Solar photooxidation of pesticides in dilute hydrogen
peroxides. J Agric Food Chem 32:231237.
Dureja P (1989) Photodecomposition of monocrotophos in soil, on plant foliage, and in
water. Bull Environ Contam Toxicol 43:239245.
Dureja P, Chattopadhyay S (1995) Photodegradation of pyrethroid insecticide flucythrinate in water and on soil surface. Toxicol Environ Chem 52:97102.
Dureja P, Johnson S (2000) Photodegradation of azadirachtin-A: a neam-based pesticide.
Curr Sci 79:17001703.
Dureja P, Mukerjee SK (1982) Photoinduced reactions: part IV. Studies on photochemical fate of 6,7,8,9,10,10-hexachloro-1,5,5a,6,9,9a-hexahydro-6,9-methano-2,4,3benzo[e]dioxathiepin-3-oxide (Endosulphan), an important insecticide. Indian J Chem
21B:411413.

Photodegradation of Pesticides

165

Dureja P, Walia S (1989) Photodecomposition of pendimetalin. Pestic Sci 25:105114.


Dureja P, Casida JE, Ruzo LO (1984) Dinitroanilines as photostabilizers for pyrethroids.
J Agric Food Chem 32:246250.
Dureja P, Walia S, Mukerjee SK (1987a) Photolysis of propiconazole. Toxicol Environ
Chem 16:6167.
Dureja P, Khazanchi R, Mukerjee SK (1987b) Photochemical decomposition of tetrachlorovinphos. Toxicol Environ Chem 15:293300.
Dureja P, Walia S, Mukerjee SK (1988) Multiphase photodegradation of quinalphos.
Pestic Sci 22:287295.
Dureja P, Walia S, Mukerjee SK (1989) Photodecomposition of isofenphos (O-ethyl
O-(2-isopropoxycarbonyl)phenyl isopropyl phosphoramidothioate). Toxicol Environ
Chem 19:187192.
Dureja P, Walia S, Prasad D (1990) Photolysis of benfuracarb. Toxicol Environ Chem
28:239244.
Eastman JA, Wesely ML, Stedman DH (1981) On the mechanisms that control vertical
ozone flux to vegetative surfaces. Proc Quad Int Ozone Symp 1:462470.
Ebing W, Schuphan I (1979) Studies on the behavior of environmental chemicals in
plants and soil quantitatively investigated in closed cultivating systems. Ecotoxicol
Environ Saf 3:133143.
Ehlers W, Letey J, Spencer WF, Farmer WJ (1969a) Lindane diffusion in soils: I. Theoretical considerations and mechanism of movement. Soil Sci Soc Am Proc 33:501
504.
Ehlers W, Spencer WF, Farmer WJ, Letey J (1969b) Lindane diffusion in soils: II. Water
content, bulk density, and temperature effects. Soil Sci Soc Am Proc 33:505508.
Elazzouzi M, Bensaoud A, Bouhaouss A, Guittonneau S, Dahchour A, Meallier P, Piccolo A (1999) Photodegradation of imazapyr in the presence of humic substances.
Fresenius Environ Bull 8:478485.
El-Nahhal Y, Nir S, Margulies L, Rubin B (1999) Reduction of photodegradation and
volatilization of herbicides in organo-clay formulation. Appl Clay Sci 14:105119.
El-Nahhal Y, Undabeytia T, Polubesova T, Mishael YG, Nir S, Rubin B (2001) Organoclay formulations of pesticides: reduced leaching and photodegradation. Appl Clay
Sci 18:309326.
El-Refai A, Hopkins TL (1966) Parathion absorption, translocation, and conversion to
paraoxon in bean plants. J Agric Food Chem 14:588592.
El-Refai A, Hopkins TL (1972) Malathion absorption, translocation, and conversion to
malaoxon in bean plants. J Assoc Offic Anal Chem 55:526531.
Emmelin C, Guittonneau S, Lamartine R, Meallier P (1993) Photodegradation of pesticides on adsorbed phases. Photodegradation of carbetamid. Chemosphere 27:757
763.
Emmelin C, Knudsen F, Guittonneau S, Meallier P (1998) Choice of silica as standardized matrix for photodegradation studies of the pesticide phenmedipham on adsorbed
phase. Fresenius Environ Bull 7:673680.
Endo S, Mintarsih TH, Kazano H (1985) Disappearance of diazinon, isoxathion and
cartap applied to rice plant. Proc Assoc Plant Prot Kyushu 31:115118.
EPA FIFRA (1999) In: Environmental fate assessment for the synthetic pyrethroids. FIFRA
Scientific Advisory Panel Meeting, Background Documents. Environmental Protection
Agency, Washington, D.C. http://www.epa.gov/scipoly/sap/1999/index.htm.

166

T. Katagi

EPA OPPTS (1993) Butyrate. In: Reregistration Eligibility Decision (RED) Butyrate.
EPA 738-R-93-014. U.S. Environmental Protection Agency, Office of Prevention,
Pesticides and Toxic Substances, Washington, DC.
EPA OPPTS (1994) Oryzalin. In: Reregistration Eligibility Decision (RED) Oryzalin.
EPA 738-R-94-016. U.S. Environmental Protection Agency, Office of Prevention,
Pesticides and Toxic Substances, Washington, DC.
EPA OPPTS (1995a) Asulam. In: Reregistration Eligibility Decision (RED) Asulam.
EPA 738-R-95-024. U.S. Environmental Protection Agency, Office of Prevention,
Pesticides and Toxic Substances, Washington, DC.
EPA OPPTS (1995b) Ethalfluralin. In: Reregistration Eligibility Decision (RED) Ethalfluralin. EPA 738-R-95-001. U.S. Environmental Protection Agency, Office of Prevention, Pesticides and Toxic Substances, Washington, DC.
EPA OPPTS (1995c) Ethion. In: Pesticide Reregistration Status (REDs, IREDS and
TREDs). U.S. Environmental Protection Agency, Washington, DC. Documents for
Ethion. Environmental fate and effects, preliminary assessment, pp 135. http://www.
epa.gov/pesticides/reregistration/status.htm
EPA OPPTS (1995d) Linuron. In: Reregistration Eligibility Decision (RED) Linuron.
EPA 738-R-95-003. U.S. Environmental Protection Agency, Office of Prevention,
Pesticides and Toxic Substances, Washington, DC.
EPA OPPTS (1996a) Amitrole. In: Reregistration Eligibility Decision (RED) Amitrole.
List A, case 0095. U.S. Environmental Protection Agency, Office of Prevention, Pesticides and Toxic Substances, Washington, DC.
EPA OPPTS (1996b) Desmedipham. In: Reregistration Eligibility Decision (RED) Desmedipham. EPA 738-R-96-004. U.S. Environmental Protection Agency, Office of
Prevention, Pesticides and Toxic Substances, Washington, DC.
EPA OPPTS (1996c) Trifluralin. In: Reregistration Eligibility Decision (RED) Desmedipham. EPA 738-R-95-040. U.S. Environmental Protection Agency, Office of Prevention, Pesticides and Toxic Substances, Washington, DC.
EPA OPPTS (1997a) Diflubenzuron. In: Reregistration Eligibility Decision (RED) Diflubenzuron. EPA 738-R-97-008. U.S. Environmental Protection Agency, Office of Prevention, Pesticides and Toxic Substances, Washington, DC.
EPA OPPTS (1997b) Propoxur. In: Reregistration Eligibility Decision (RED) Propoxur.
EPA 738-R-97-009. U.S. Environmental Protection Agency, Office of Prevention,
Pesticides and Toxic Substances, Washington, DC.
EPA OPPTS (1998a) Azinphos-methyl. In: Pesticide Reregistration Status (REDs, IREDS
and TREDs). U.S. Environmental Protection Agency, Washington, D.C. Documents
for Azinphos-methyl. Environmental fate and effects, preliminary assessment, pp 1
162. http://www.epa.gov/pesticides/reregistration/status.htm.
EPA OPPTS (1998b) Butralin. In: Reregistration Eligibility Decision (RED) Butralin.
EPA 738-R-97-09. U.S. Environmental Protection Agency, Office of Prevention, Pesticides and Toxic Substances, Washington, DC.
EPA OPPTS (1998c) Dichlobenil. In: Reregistration Eligibility Decision (RED) Dichlobenil. EPA 738-R-98-003. U.S. Environmental Protection Agency, Office of Prevention, Pesticides and Toxic Substances, Washington, DC.
EPA OPPTS (1998d) Dicofol. In: Reregistration Eligibility Decision (RED) Dicofol.
EPA 738-R-98-018. U.S. Environmental Protection Agency, Office of Prevention,
Pesticides and Toxic Substances, Washington, DC.

Photodegradation of Pesticides

167

EPA OPPTS (1998e) Iprodione. In: Reregistration Eligibility Decision (RED) Iprodione.
EPA 738-R-98-019. U.S. Environmental Protection Agency, Office of Prevention,
Pesticides and Toxic Substances, Washington, DC.
EPA OPPTS (1998f) Isofenphos. In: Pesticide Reregistration Status (REDs, IREDS and
TREDs). U.S. Environmental Protection Agency, Washington, DC. Documents for
Isofenphos. Environmental fate and effects, preliminary assessment, pp 147. http://
www.epa.gov/pesticides/reregistration/status.htm.
EPA OPPTS (1998g) Metribuzin. In: Reregistration Eligibility Decision (RED) Metribuzin. EPA 738-R-97-006. U.S. Environmental Protection Agency, Office of Prevention, Pesticides and Toxic Substances, Washington, DC.
EPA OPPTS (1998h) Terbacil. In: Reregistration Eligibility Decision (RED) Terbacil.
EPA 738-R-97-011. U.S. Environmental Protection Agency, Office of Prevention,
Pesticides and Toxic Substances, Washington, DC.
EPA OPPTS (1999a) Bensulide. In: Pesticide Reregistration Status (REDs, IREDS and
TREDs). U.S. Environmental Protection Agency, Washington, DC. Documents for
Bensulide. Revised environmental fate and effects assessment, pp 1102. http://www.
epa.gov/pesticides/reregistration/status.htm.
EPA OPPTS (1999b) Captan. In: Reregistration Eligibility Decision (RED) Captan. EPA
738-R-99-015. U.S. Environmental Protection Agency, Office of Prevention, Pesticides and Toxic Substances, Washington, DC.
EPA OPPTS (1999c) Chlorothalonil. In: Reregistration Eligibility Decision (RED) Chlorothalonil. EPA 738-R-99-004. U.S. Environmental Protection Agency, Office of Prevention, Pesticides and Toxic Substances, Washington, DC.
EPA OPPTS (1999d) Disulfoton. In: Pesticide Reregistration Status (REDs, IREDS and
TREDs). U.S. Environmental Protection Agency, Washington, DC. Documents for
Disulfoton. Revised environmental fate and effects assessment, pp 1102. http://www.
epa.gov/pesticides/reregistration/status.htm.
EPA OPPTS (1999e) Methyl parathion. In: Pesticide Reregistration Status (REDs, IREDS
and TREDs). U.S. Environmental Protection Agency, Washington, DC. Documents
for Methyl Parathion. Revised environmental fate and effects assessment, pp 187.
http://www.epa.gov/pesticides/reregistration/status.htm.
EPA OPPTS (2000) Diazinon. In: Pesticide Reregistration Status (REDs, IREDS and
TREDs). U.S. Environmental Protection Agency, Washington, DC. Documents for Diazinon. Revised science chapter, pp 1171. http://www.epa.gov/pesticides/reregistration/
status.htm.
EPA OPPTS (2001) Thiophanate-methyl. In: Pesticide Reregistration Status (REDs, IREDS
and TREDs). U.S. Environmental Protection Agency, Washington, DC. Documents
for Thiophanate-methyl. Preliminary risk Assessments, pp 1124. http://www.epa.
gov/pesticides/reregistration/status.htm.
Fang CH (1977) Effects of soils on the degradation of herbicide alachlor under the light.
J Chin Agric Chem Soc 15:5359.
Farmer WJ, Jensen CR (1970) Diffusion and analysis of carbon-14 labeled dieldrin in
soils. Soil Sci Soc Am Proc 34:2831.
Feely WF, Crouch LS, Arison BH, VandenHeuvel WJA, Colwell LF, Wislocki PG
(1992) Photodegradation of 4-(epi-methylamino)-4-deoxyavermectin B1a thin films
on glass. J Agric Food Chem 40:691696.
Fenet H, Beltran E, Gadji B, Cooper JF, Coste CM (2001) Fate of a phenylpyrazole in
vegetation and soil under tropical field conditions. J Agric Food Chem 49:12931297.

168

T. Katagi

Feng W, Nansheng D (2000) Photochemistry of hydrolytic iron (III) species and photoinduced degradation of organic compounds. A minireview. Chemosphere 41:1137
1147.
Fischer AM, Kliger DS, Winterle JS, Mill T (1985) Direct observation of phototransients
in natural waters. Chemosphere 14:12991306.
Fischer AM, Winterle JS, Mill T (1987) Primary photochemical processes in photolysis
mediated by humic substances. In: Zika RG, Cooper WJ (eds) Photochemistry of
Environmental Aquatic Systems. ACS Symposium Series 327. American Chemical
Society, Washington, DC, pp 141156.
Fleeker JR, Lacy HM (1977) Photolysis of methyl 2-benzimidaolecarbamate. J Agric
Food Chem 25:5155.
Floer-Muller H, Schwack W (2001) Photochemistry of organophosphorus insecticides.
Rev Environ Contam Toxicol 172:129228.
Fontan J, Minga A, Lopez A, Druilhet A (1992) Vertical ozone profiles in a pine forest.
Atmos Environ 26A:863869.
Foote CS (1968a) Mechanism of photosensitized oxidation. Science 162:963970.
Foote CS (1968b) Photosensitized oxygenations and the role of singlet oxygen. Acc
Chem Res 1:104110.
Fowler D, Flechard C, Cape JN, Storeton-West RL, Coyle M (2001) Measurements of
ozone deposition to vegetation quantifying the flux, the stomatal and non-stomatal
components. Water Air Soil Pollut 130:6374.
Frank MP, Graebing PG, Chib JS (2002) Effect of soil moisture and sample depth on
pesticide photolysis. J Agric Food Chem 50:26072614.
Frei RW, Nomura NS (1968) A comparison of new techniques for the detection and
quantitative determination of triazine herbicides separated by thin-layer chromatography. II. Reflectance spectroscopy. Mikrochim Acta (Wien) 3:565573.
Frimmel FH (1994) Photochemical aspects related to humic substances. Environ Int 20:
373385.
Fruekilde P, Hjorth J, Jensen NR, Kotzias D, Larson B (1998) Ozonolysis at vegetation
surfaces: a source of acetone, 4-oxopentanal, 6-methyl-5-hepten-2-one, and geranyl
acetone in the troposphere. Atmos Environ 32:18931902.
Fujii Y, Asaka S, Misato T (1979) Photodegradation of dipropyl 4-(methylthio)phenyl
phosphate (propaphos, Kayaphos). J Pestic Sci 4:361366.
Fukushima M, Tatsumi K (1999) Photocatalytic reaction by iron (III)-humate complex
and its effect on the removal of organic pollutant. Toxicol Environ Chem 73:103
116.
Fukushima M, Tatsumi K (2001) Degradation pathways of pentachlorophenol by photoFenton systems in the presence of iron (III), humic acid, and hydrogen peroxide.
Environ Sci Technol 35:17711778.
Fukushima M, Tatsumi K, Morimoto K (2000) The fate of aniline after a photo-Fenton
reaction in an aqueous system containing iron (III), humic acid, and hydrogen peroxide. Environ Sci Technol 34:20062013.
Fukushima M, Tatsumi K, Nagao S (2001) Degradation characteristics of humic acid
during photo-Fenton processes. Environ Sci Technol 35:36833690.
Fukushima M, Fujisawa T, Katagi T, Takimoto Y (2003) Metabolism of fenitrothion and
conjugation of 3-methyl-4-nitrophenol in tomato plant (Lycopersicon esculentum).
J Agric Food Chem 51:50165023.
Gab S, Saravanja V, Korte F (1975a) Irradiation studies of aldrin and chlordane adsorbed
on a silica gel surfaces. Bull Environ Contam Toxicol 13:301306.

Photodegradation of Pesticides

169

Gab S, Nitz S, Parlar H, Korte F (1975b) Photomineralization of certain aromatic xenobiotica. Chemosphere 4:251256.
Garau V, Angioni A, Real AAD, Russo M, Cabras P (2002) Disappearance of azoxystrobin, pyrimethanil, cyprodinil and fludioxinil on tomatoes in a greenhouse. J Agric
Food Chem 50:19291932.
Gates DM, Keegan HJ, Schleter JC, Weidner VR (1965) Spectral properties of plants.
Appl Opt 4:1120.
Gerecke AC, Canonica S, Muller SR, Scharer M, Schwarzenbach RP (2001) Quantification of dissolved natural organic matter (DOM) mediated phototransformation of phenylurea herbicides in lakes. Environ Sci Technol 35:39153923.
Gerstl Z, Yaron B, Nye PH (1979) Diffusion of a biodegradable pesticide: I. In a biologically inactive soil. Soil Sci Soc Am J 43:839842.
Ghosal DN, Mukherjee SK (1972) A spectrophotometric study of dye aggregation on
clay surfaces. Indian J Chem 10:835837.
Gil Garcia MD, Martinez Vidal JL, Martinez Galera M, Rodriguez Torreblanca C, Gonzalez C (1997) Determination and degradation of methomyl in tomatoes and green
beans grown in greenhouses. J Assoc Offic Anal Chem Int 80:633638.
Glass B (1975) Photosensitization and luminescence of picloram. J Agric Food Chem
23:11091112.
Goetz AJ, Lavy TL, Gbur Jr EE (1990) Degradation and field persistence of imazethapyr.
Weed Sci 38:421428.
Gohre K, Miller GC (1983) Singlet oxygen generation on soil surfaces. J Agric Food
Chem 31:11041108.
Gohre K, Miller GC (1985) Photochemical generation of singlet oxygen on non-transition-metal oxide surfaces. J Chem Soc Faraday Trans 1 81:793800.
Gohre K, Miller GC (1986) Photooxidation of thioether pesticides on soil surfaces.
J Agric Food Chem 34:709713.
Gohre K, Scholl R, Miller GC (1986) Singlet oxygen reaction on irradiated soil surfaces.
Environ Sci Technol 20:934938.
Gong A, Ye C, Wang X, Lei Z, Liu J (2001) Dynamics and mechanism of ultraviolet
photolysis of atrazine on soil surface. Pest Manag Sci 57:380385.
Gould IR (1989a) Conventional light sources. In: Scaiano JC (ed) CRC Handbook of
Organic Photochemistry, vol 1. CRC Press, Boca Raton, FL, pp 317.
Gould IR (1989b) Conventional light sources. In: Scaiano JC (ed) CRC Handbook of
Organic Photochemistry, vol 1. CRC Press, Boca Raton, FL, pp 155196.
Graebing P, Frank M, Chib JS (2002) Effects of fertilizer and soil components on pesticide photolysis. J Agric Food Chem 50:73327339.
Graebing P, Frank M, Chib JS (2003) Soil photolysis of herbicides in a moisture- and
temperature-controlled environment. J Agric Food Chem 51:43314337.
Graham-Bryce IJ (1969) Diffusion of organophosphorus insecticides in soils. J Sci Food
Agric 20:489494.
Grayson BT, Williams KS, Freehauf PA, Pease RR, Ziesel WT, Sereno RL, Reinsfelder
RE (1987) The physical and chemical properties of the herbicide cinmethylin (SD
95481). Pestic Sci 21:143153.
Greenhalgh R, Marshall WD (1976) Ultraviolet irradiation of fenitrothion and the synthesis of the photolytic oxidation products. J Agric Food Chem 24:708713.
Griesbeck AG, Gorner H (1999) Laser flash photolysis study of N-alkylated phthalimides. J Photochem Photobiol A Chem 129:111119.

170

T. Katagi

Gunther FA (1969) Insecticide residues in California citrus fruits and products. Residue
Rev 28:1119.
Gunther FA, Ott DE, Ittig M (1970) The oxidation of prathion to paraoxon. II. By use
of ozone. Bull Environ Contam Toxicol 5:8794.
Gustafson DI, Holden LR (1990) Nonlinear pesticide dissipation in soil: a new model
based on spatial variability. Environ Sci Technol 24:10321038.
Guth JA (1981) Experimental approaches to studying the fate of pesticides in soil. In:
Hutson DH, Roberts TR (eds) Progress in Pesticide Biochemistry, vol 1. Wiley, New
York, pp 85114.
Haag WR, Hoigne J (1986) Singlet oxygen in surface waters. 3. Photochemical formation
and steady-state concentrations in various types of waters. Environ Sci Technol 20:
341348.
Haas K, Schonherr J (1979) Composition of soluble cuticular lipids and water permeability of cuticular membranes from citrus leaves. Planta (Berl) 146:399403.
Hainzl D, Casida JE (1996) Fipronil insecticide: novel photochemical desulfinylation
with retention of neurotoxocity. Proc Natl Acad Sci USA 93:1276412767.
Halder P, Barua AS, Raha P, Biswas B, Pal S, Bhattacharys A, Bedi S, Chowdhury A
(1989) Studies on the photodegradation of pendimethalin in solutions and in Kalyani
soil. Chemosphere 18:16111619.
Harrison RB, Holmes DC, Roburn J, Tatton JOG (1967) The fate of some organochlorine
pesticides on leaves. J Sci Food Agric 18:1015.
Harrison SK, Thomas SM (1990) Interaction of surfactants and reaction media on photolysis of chlorimuron and metsulfuron. Weed Sci 38:620624.
Harrison SK, Wax LM (1985) The effect of adjuvants and oil carriers on photodecomposition of 2,4-D, bentazon and haloxyfop. Weed Sci 34:8187.
Harvey J Jr, Han JCY, Reiser RW (1978) Metbolism of oxamyl in plants. J Agric Food
Chem 26:529536.
Hautala RR (1978) Surfactant effects on pesticide photochemistry in water and soil.
EPA Technical Report Data. EPA-600/3-78-060 (PB-285175). U.S. Environmental
Protection Agency, Washington, DC, pp 183.
Hazen JL (2000) Adjuvants: terminology, classification, and chemistry. Weed Technol
14:773784.
Herbert VR, Miller GC (1990) Depth dependence of direct and indirect photolysis on
soil surfaces. J Agric Food Chem 38:913918.
Herrmann M, Kotzias D, Korte F (1985) Photochemical behavior of chlorsulfuron in
water and in adsorbed phase. Chemosphere 14:38.
Hess FD, Foy CL (2000) Interaction of surfactants with plant cuticles. Weed Technol
14:807813.
Hirahara Y, Ueno H, Nakamuro K (2001) Comparative photodegradation study of fenthion and disulfoton under irradiation of different light sources in liquid- and solidphases. J Health Sci 47:129135.
Hirahara Y, Okuno T, Ueno H, Nakamuro K (2003) Photooxidation mechanism of fenthion by singlet oxygen: evidence by ESR analysis with a selective spin trapping
agent. J Health Sci 49:3439.
Hirayama Y, Sayato Y, Nakamuro K (1998) Studies on photochemical behaviors of
pesticides in environment. Jpn J Toxicol Environ Health 44:451461.
Hirt RC, Schmitt RG, Searle ND, Sullivan AP (1960) Ultraviolet spectral energy distribution of natural sunlight and accelerated test light sources. J Opt Soc Am 50:706
713.

Photodegradation of Pesticides

171

Hoigne J, Faust BC, Haag WR, Scully FE Jr, Zepp RG (1989) Aquatic humic substances
as sources and sinks of photochemically produced transient reactants. In: Suffet IH,
MacCarthy PM (eds) Aquatic Humic Substances: Influence on Fate and Treatment of
Pollutants. Advances in Chemistry Series 219. American Chemical Society, Washington, DC, pp 363381.
Holland F, Hessling M, Hofzumahaus A (1995) In situ measurement of tropospheric OH
radicals by laser-induced fluorescence: a description of the KFA instrument. J Atmos
Sci 52:33933401.
Hollrigl-Rosta A, Kreuzig R, Bahadir M (1999) Investigation on the metabolic fate of
prochloraz in soil under field and laboratory conditions. Pestic Sci 55:531538.
Holmes MG, Keiller DR (2002) Effects of pubescence and waxes on the reflectance of
leaves in the ultraviolet and photosynthetic wavebands: a comparison of a range of
species. Plant Cell Environ 25:8593.
Holmstead RL, Casida JE, Ruzo LO, Fullmer DG (1978a) Pyrethroid photodecomposition: permethrin. J Agric Food Chem 26:590595.
Holmstead RL, Fullmer DG, Ruzo LO (1978b) Pyrethroid photodecomposition: pydrin.
J Agric Food Chem 26:954959.
Horn DHS, Lamberton JA (1962) Long-chain -diketones from plant waxes. Chem Ind
(Lond) 20362037.
Hosokawa S, Miyamoto J (1974) Metabolism of 14C-labeled sumithion, O,O-dimethyl O(3-methyl-4-nitrophenyl) phosphorothioate in apples. Botyu-Kagaku 39:4953.
Hubbs CW, Lavy TL (1990) Dissipation of norflurazon and other persistent herbicides
in soil. Weed Sci 38:8188.
Huling SG, Arnold RG, Sierka RA, Miller MR (1998) Measurement of hydroxyl radical
activity in a soil slurry using the spin trap -(4-pyridyl-1-oxide)-N-tert-butylnitrone.
Environ Sci Technol 32:34363441.
Hulpke H, Stegh R, Wilmes R (1983) Light-induced transformations of pesticides on
silica gel as a model system for photodegradation on soil. In: Miyamoto J, Kearney
PC (eds) Human Welfare and the Environment. Proceedings 5th International Congress on Pesticide Chemistry, vol 3. Pergamon Press, Oxford, pp 323326.
Hustert K, Moza PN (1994) Photocatalytic degradation of azo dyes by semiconducting
iron compounds. Fresenius Environ Bull 3:762767.
Hustert K, Moza PN (1997) Photochemical degaradtion of dicarboxiimide fungicides in
the presence of soil constituents. Chemosphere 35:3337.
Hustert K, Moza PN, Kettrup A (1999) Photochemical degradation of carboxin and oxycarboxin in the presence of humic substances and soil. Chemosphere 38:34233429.
Ishikawa K, Nakamura Y, Niki Y, Kuwatsuka S (1977) Photodegradation of benthiocarb
herbicide. J Pestic Sci 2:1725.
Ishizuka K, Takase I, Ei-Tan K, Mitsui S (1973) Absorption and translocation of Oethyl S,S-diphenyl phosphorodithioate (Hinosan) in rice plants. Agric Biol Chem 37:
13071316.
Isobe N, Matsuo M, Miyamoto J (1984) Novel photoproducts of allethrin. Tetrahedron
Lett 25:861864.
Ivie GW, Bull DL (1976) Photodegradation of O-ethyl O-[4-(methylthio)phenyl] S-propyl phosphorodithioate (BAY NTN9306). J Agric Food Chem 24:10531057.
Ivie GW, Casida JE (1970) Enhancement of photoalteration of cyclodiene insecticide
chemical residues by rotenone. Science 167:16201622.

172

T. Katagi

Ivie GW, Casida JE (1971a) Sensitized photodecomposition and photosensitizer activity


of pesticide chemicals exposed to sunlight on silica gel chromatoplates. J Agric Food
Chem 19:405409.
Ivie GW, Casida JE (1971b) Photosensitizers for accelerated degradation of chlorinated
cyclodienes and other insecticide chemicals exposed to sunlight on bean leaves.
J Agric Food Chem 19:410416.
Ivie GW, Knox JR, Khalifa S, Yamamoto I, Casida JE (1972) Novel photoproducts of
heptachlor epoxide, trans-chlordane and trans-nonachlor. Bull Environ Contam Toxicol 76:376382.
Jacques GL, Harvey RG (1979) Adsorption and diffusion of dinitroaniline herbicides in
soils. Weed Sci 27:450455.
Jacob TA, Carlin JR, Walker RW, Wolf FJ, Vanden Heuvel JA (1975) Photolysis of
thiobendazole. J Agric Food Chem 23:704709.
Jacques GL, Harvey RG (1979) Adsorption and diffusion of dinitroaniline herbicides in
soils. Weed Sci 27:450455.
Jaffe HH, Orchin M (1962) Theory and Applications of Ultraviolet Spectroscopy. Wiley,
Sons, New York.
Jahn C, Zorn H, Petersen A, Schwack W (1999) Structure-specific detection of plant
cuticle bound residues of chlorothalonil by ELISA. Pestic Sci 55:11671176.
Jensen-Korte U, Anderson C Spiteller M (1987) Photodegradation of pesticides in the
presence of humic substances. Sci Total Environ 62:335340.
Jernberg KM, Lee PW (1999) Fate of famoxadone in the environment. Pestic Sci 55:
587589.
Jirkovsky J, Faure V, Boule P (1997) Photolysis of diuron. Pestic Sci 50:4252.
Joiner RL, Baetcke KP (1973) Parathion: persistence on cotton and identification of its
photoalteration products. J Agric Food Chem 21:391396.
Johnston LJ, DeMayo P, Wong SK (1984) Surface photochemistry: decomposition of
azobis(isobutyronitrile) on dry silica gel. J Org Chem 49:2026.
Kanofsky JR (2000) Assay for singlet oxygen generation by peroxidases using 1270-nm
chemiluminescence. In: Packer L, Sies H (eds) Singlet Oxygen, UV-A, and Ozone.
Methods in Enzymology, vol 319. Academic Press, New York, pp 5967.
Kanofsky JR, Sima PD (2000) Assay for singlet oxygen generation by plant leaves exposed to ozone. In: Packer L, Sies H (eds) Singlet Oxygen, UV-A, and Ozone. Methods in Enzymology, vol 319. Academic Press, New York, pp 512520.
Katagi T (1989) Molecular orbital approaches to the photolysis of organophosphorus
insecticide fenitrothion. J Agric Food Chem 37:11241130.
Katagi T (1990) Photoinduced oxidation of the organophosphorus fungicide tolclofosmethyl on clay minerals. J Agric Food Chem 38:15951600.
Katagi T (1991) Photodegradation of the pyrethroid insecticide esfenvalerate on soil,
clay minerals, and humic acid surfaces. J Agric Food Chem 39:13511356.
Katagi T (1992) Photodegradation of 3-phenoxybenzoic acid in water and on solid surfaces. J Agric Food Chem 40:12691274.
Katagi T (1993a) Photochemistry of organophosphorus herbicide butamifos. J Agric
Food Chem 41:496501.
Katagi T (1993b) Effect of moisture content and UV irradiation on degradation of fenpropathrin on soil surfaces. J Pestic Sci 18:333341.
Katagi T (2002a) Experimental and theoretical studies on photodegradation of fungicide
diniconazole. J Pestic Sci 27:111117.

Photodegradation of Pesticides

173

Katagi T (2002b) Abiotic hydrolysis of pesticides in the aquatic environment. Rev Environ Contam Toxicol 175:79261.
Katagi T, Mikami N (2000) Primary metabolism of agrochemicals in plants. In: Roberts
T (ed) Metabolism of Agrochemicals in Plants. Wiley, New York, pp 43106.
Katagi T, Kikuzono Y, Mikami N, Matsuda T, Miyamoto J (1988) A theoretical approach to photochemistry of pyrethroids possessing the cyclopropane ring. J Pestic
Sci 13:129132.
Ke TS, Hsieh YM, Tsai LS, Wang Y (1940) The ultraviolet absorption spectrum of
rotenone. J Chin Chem Soc 6:4043.
Kerstiens G, Lendzion K (1989) Interactions between ozone and plant cuticles. I. Ozone
deposition and permeability. New Phytol 112:1319.
Ketchersid ML, Merkle MG (1975) Persistence and movement of perfluidone in soil.
Weed Sci 23:344348.
Keum YS, Liu KH, Seo JS, Kim JH, Kim K, Kim YH, Kim PJ (2002) Dissipation of
foliar residues of diafenthiuron and its metabolites. Bull Environ Contam Toxicol 68:
845851.
Kieatiwong S, Miller GC (1992) Photolysis of aryl ketones with varying vapor pressures
on soil. Environ Toxicol Chem 11:173179.
Kieatiwong S, Nguyen LV, Herbert VR, Hackett M, Miller GC, Miille MJ, Mitzel R
(1990) Photolysis of chlorinated dioxins in organic solvents and on soils. Environ Sci
Technol 24:15751580.
Kieber DJ, Blough NV (1990) Determination of carbon-centered radicals in aqueous
solution by liquid chromatography with fluorescence detection. Anal Chem 62:2275
2283.
Kimmel EC, Casida JE, Ruzo LO (1982) Identification of mutagenic photoproducts of
the pyrethroids allethrin and terallethrin. J Agric Food Chem 30:623626.
Kitchener JA (1946) The photochemistry of solids. Sci J R Colloid Sci 19:114.
Klehr M, Iwan J, Riemann J (1983) An experimental approach to the photolysis of pesticides adsorbed on soil: thiadiazuron. Pestic Sci 14:359366.
Kleier DA (1994) Environmental effects on the photodegradation of pesticides. In: Comparing Glasshouse & Field Pesticide Performance, vol II. BCPC Monograph No. 59.
The Brighton Crop Protection Council, Farnham, Surrey (UK), pp 97109.
Klein W, Kohli J, Weisgerber I, Korte F (1973) Fate of aldrin-14C in potatoes and soil
under outdoor conditions. J Agric Food Chem 21:152156.
Knowles CO, Sen Gupta AK (1969) Photodecomposition of the acaricide N-(4-chloroo-tolyl)-N,N-dimethylformamidine. J Econ Entomol 62:344348.
Kofler M, Langemann A, Ruegg R, Chopard-dit-Jean LH, Rayroud A, Isler O (1959)
Structure of a plant quinone with isopremoid side chain. Helv Chim Acta 42:1283
1292.
Kolb CA, Kazer MA, Kopecky J, Zotz G, Riederer M, Pfundel EE (2001) Effects of
natural intensities of visible and ultraviolet radiation on epidermal ultraviolet screening and photosynthesis in grape leaves. Plant Physiol 127:863875.
Konstantinou IK, Zarkadis AK, Albanis TA (2001) Photodegradation of selected herbicides in various natural waters and soils under environmental conditions. J Environ
Qual 30:121130.
Kopf G, Schwack W (1995) Photodegradation of the carbamate insecticide ethiofencarb.
Pestic Sci 43:303309.

174

T. Katagi

Koshy KT, Burdick MD, Knuth DW (1983) Multiphase photodegradation of methyl N[[[[[(1,1-dimethylethyl) (5,5-dimethyl-2-thioxo-1,3,2-dioxaphosphorinan-2-yl)-amino]thio]
methylamino] carbonyl]oxy]ethanimidothiolate. J Agric Food Chem 31:625629.
Kotzias D, Herrmann M, Zsolnay A, Russi H, Korte F (1986) Photochemical reactivity
of humic materials. Naturwissenschaften 73:3536.
Krause RT (1983) Determination of fluorescent pesticides and metabolites by reversedphase high-performance liquid chromatography. J Chromatogr 255:497510.
Krieger MS, Yoder RN, Gibson R (2000) Photolytic degradation of florasulam on soil
and in water. J Agric Food Chem 48:37103717.
Kromer T, Ophoff H, Fuhr F, Stork A (1999) Photodegradation and volatilization of
parathion-methyl on glass and soil dust under laboratory conditions. In: Human Environmental Exposure to Xenobiotics. Proceedings, 11th Symposium on Pesticide
Chemistry. Goliardica Pavese, Pavia, Italy, pp 363374.
Kulshrestha G, Mukerjee SK (1986) The photochemical decomposition of the herbicide
isoproturon. Pestic Sci 17:489494.
Kumar Y, Semeluk GP, Silk PJ, Unger I (1974) The photochemistry of carbamates. VI:
The photodecomposition of meobal (3,4-xylyl-N-methylcarbamate) and mesurol
(4-methylthio-3,5-xylyl-N-methylcarbamate). Chemosphere 3:2327.
Lackhoff M, Niessner R (2002) Photocatalytic atrazine degradation by synthetic minerals, atmospheric aerosols, and soil particles. Environ Sci Technol 36:53425347.
Larson RA, Schlauch MB, Marley KA (1991) Ferric ion promoted photodecomposition
of triazines. J Agric Food Chem 39:20572062.
Larson RA, Marley KA (1999) Singlet oxygen in the environment. In: Boule P (ed) The
Handbook of Environmental Chemistry, vol 2, part L. Springer-Verlag, Berlin, pp
123137.
Larsson S, Svenningsson M (1986) Cuticular transpiration and epicuticular lipids of primary leaves of barley (Hordeum vulgare). Physiol Plant 68:1319.
Lee JSK, Huang PM (1995) Photochemical effects on the abiotic transformations of
polyphenolics as catalyzed by Mn(II) oxide. In: Huang PM, Berthelin J, Bollag JM,
McGill WB (eds) Environmental Impacts of Soil Component Interactions: Land Quality, Natural and Anthropogenic Organics, vol 1. Lewis, Boca Raton, FL, pp 177189.
Lee PW, Srearns SM, Powell WR (1988) Metabolic fate of fenvalerate in wheat plants.
J Agric Food Chem 36:189193.
Lee PW, Stearns SM, Hernandez H, Powell WR, Naidu MV (1989) Fate of dicrotophos
in the soil environment. J Agric Food Chem 37:11691174.
Lee PW, Fukuto JM, Hernandez H, Stearns SM (1990) Fate of monocrotophos in the
environment. J Agric Food Chem 38:567573.
Leermakers PA, Thomas HT, Weis LD, James FC (1966) Spectra and photochemistry
of molecules adsorbed on silica gel. IV. J Am Chem Soc 88:50755083.
Leifer A (1988) The Kinetics of Environmental Aquatic Photochemistry. Theory and
Practice. ACS Professional Reference Book. American Chemical Society, Washington, DC.
Lendzian KJ, Kerstiens G (1991) Sorption and transport of gases and vapors in plant
cuticles. Rev Environ Contam Toxicol 121:65128.
Leuch DB, Bowman MC (1968) Residues of fenthion and five of its metabolites: their
persistence in corn and grape forages. J Econ Entomol 61:15941597.
Lewis RG (1976) Sampling and analysis of airborne pesticides. In: Lee RE Jr (ed) Air
Pollution from Pesticides and Agricultural Processes. CRC Press, Cleveland, OH, pp
5194.

Photodegradation of Pesticides

175

Liang TT, Lichtenstein EP (1976) Effects of soils and leaf surfaces on the photodecomposition of [14C]Azinphosmethyl. J Agric Food Chem 24:12051210.
Lichtenstein EP, Fuhrmann TW, Schulz KR, Liang TT (1973) Effects of field application
methods on the persistence and metabolism of phorate in soils and its translocation
into crops. J Econ Entomol 66:863866.
Linders J, Mensink H, Stephenson G, Wauchope D, Racke K (2000) Foliar interception
and retention values after pesticide application. A proposal for standardized values
for environmental risk assessment. Pure Appl Chem 72:21992218.
Lindquist DA, Bull DL (1967) Fate of 3-hydroxy-N-methyl-cis-crotonamide dimethyl
phosphate in cotton plants. J Agric Food Chem 15:267272.
Liu PY, Zheng MH, Xu XB (2002) Phototransformation of polychlorinated dibenzop-dioxins from photolysis of pentachlorophenol on soil surface. Chemosphere 46:
11911193.
Liu X, Iu KK, Mao Y, Thomas JK (1994) Photoinduced reactions on clay and model
surfaces. In: Helz GR, Zepp RG, Crosby DG (eds) Aquatic Surface Photochemistry.
Lewis, Boca Raton, FL, pp 187195.
Logan JA (1999) An analysis of ozonesonde data for the troposphere: recommendations
for testing 3-D models and development of a gridded climatology for tropospheric
ozone. J Geophys Res 104:1611516149.
Lucier GW, Menzer RE (1968) Metabolism of dimethoate in bean plants in relation to
its mode of application. J Agric Food Chem 16:936945.
Lucier GW, Menzer RE (1970) Nature of oxidative metabolites of dimethoate formed in
rats, liver microsomes, and bean plants. J Agric Food Chem 18:698704.
Maguire RJ (1990) Chemical and photochemical isomerization of deltamethrin. J Agric
Food Chem 38:16131617.
Mahnken GE, Weber JB (1988) Capillary movement of triasulfuron and chlorsulfuron in
Rion sandy loam soil. Proc South Weed Sci Soc 41:332336.
Makary MH, Riskallah MR, Hegazy ME, Belal MH (1981) Photolysis of phoxim on
glass and on tomato leaves. Bull Environ Contam Toxicol 26:413419.
Mallet V, Surette DP (1974) Fluorescence of pesticides by treatment with heat, acid or
base. J Chromatogr 95:243246.
Mamouni A, Schmitt P, Mansour M, Schiavon M (1992) Abiotic degradation pathways
of isoxaben in the environment. Pestic Sci 35:1320.
Manahan SE (1994) The geosphere and geochemistry. In: Manahan SE (ed) Environmental Chemistry, 6th Ed. Lewis, Boca Raton, FL, pp 433456.
Mansager ER, Still GG, Frear DS (1979) Fate of [14C]diflubenzuron on cotton and in
soil. Pestic Biochem Physiol 12:172182.
Mansour M, Feicht EA, Behechti A, Scheunert I (1997) Experimental approaches to
studying the photostability of selected pesticides in water and soil. Chemosphere 35:
3950.
Mao Y, Thomas JK (1993) Photoinduced electron transfer and subsequent chemical reactions of adsorbed thianthrene on clay surfaces. J Org Chem 58:66416649.
Marcheterre L, Choudhry GG, Webster GRB (1988) Environmental photochemistry of
herbicides. Rev Environ Contam Toxicol 103:61126.
Margulies L, Rozen H, Cohen E (1985) Energy transfer at the surface of clays and
protection of pesticides from photodegradation. Nature (Lond) 315:658659.
Margulies L, Cohen E, Rozen H (1987) Photostabilization of bioresmethrin by organic
cations on a clay surface. Pestic Sci 18:7987.

176

T. Katagi

Margulies L, Rozen H, Cohen E (1988) Photostabilization of a nitromethylene heterocycle insecticide on the surface of montmorillonite. Clays Clay Miner 36:159164.
Margulies L, Stern T, Rubin B, Ruzo LO (1992) Photostabilization of trifluralin adsorbed
on a clay matrix. J Agric Food Chem 40:152155.
Margulies L, Rozen H, Stern T, Rytwo G, Rubin B, Ruzo LO, Nir S, Cohen E (1993)
Photostabilization of pesticides by clays and chromophores. Arch Insect Biochem
Physiol 22:467486.
Martnez Galera M, Martnez Vidal JL, Egea Gonzalez FJ, Gil Garca MD (1997) A
study of fenpropathrin residues in tomatoes and green beans grown in greenhouses in
Spain. Pestic Sci 50:127134.
Martnez Vidal JL, Egea Gonzalez FJ, Martnez Galera M, Castro Cano ML (1998)
Diminution of chlorpyrifos and chlorpyrifos oxon in tomatoes and green beans grown
in greenhouses. J Agric Food Chem 46:14401444.
Matsuo H, Casida JE (1970) Photodegradation of two dinitrophenolic pesticide chemicals, dinobuton and dinoseb, applied to bean leaves. Bull Environ Contam Toxicol 5:
7278.
Mazellier P, Bolte M (2000) Heterogeneous light-induced transformation of 2,6-dimethylphenol in aqueous suspensions containing goethite. J Photochem Photobiol A Chem
132:129135.
McDonald RE, Nordby HE, McCollum TG (1993) Epicuticular wax morphology and
composition are related to grapefruit chilling injury. Hortic Sci 28:311312.
McFarlane JC (1995) Anatomy and physiology of plant conductive systems. In: Trapp
S, McFarlane JC (eds) Plant Contamination: Modeling and Simulation of Organic
Chemical Processes. Lewis, Boca Raton, FL, pp 1334.
McPhail DB, Hartley RD, Gardner PT, Duthie GG (2003) Kinetic and stoichiometric
assessment of the antioxidant activity of flavonoids by electron spin resonance spectroscopy. J Agric Food Chem 51:16841690.
Meallier P (1999) Phototransformation of pesticides in aqueous solution. In: The Handbook of Environmental Chemistry, vol 2, part L. Springer-Verlag, Berlin, pp 241
261.
Megahed HS, Steurbaut W, Dejonckheere W (1987) Influence of the presence of soybean
oilsurfactant combinations on the rainfastness and the photodegradation of insecticide deposits. Meded Fac Landbouwwet Rijkuniv Gent 52:713719.
Meikle RW, Kurihara NH, DeVries DH (1983) Chlorpyrifos: the photodecomposition
rates in dilute aqueous solution and on a surface, and the volatilization rates from a
surface. Arch Environ Contam Toxicol 12:189193.
Mikami N, Ohkawa H, Miyamoto J (1976) Photodecomposition of surecide (O-ethyl
O-4-cyanophenyl phenylphosphonothioate) and cyanox (O,O-dimethyl O-4-cyanophenyl phosphorothioate). J Pestic Sci 1:273281.
Mikami N, Ohkawa H, Miyamoto J (1977a) Stereospecificity in oxidation of the optical
isomers of O-ethyl O-2-nitro-5-methylphenyl N-isopropyl phosphoramidothioate
(S-2571) by liver mixed function oxidase and UV light. J Pestic Sci 2:119126.
Mikami N, Ohkawa H, Miyamoto J (1977b) Photodecomposition of salithion (2-methoxy4H-1,3,2-benzodioxaphosphorin-2-sulfide) and phenthoate (O,O-dimethyl S--ethoxycarbonylbenzyl phosphorodithioate). J Pestic Sci 2:279290.
Mikami N, Takahashi N, Hayashi K, Miyamoto J (1980) Photodegradation of fenvalerate
(Sumicidin) in water and on soil surfaces. J Pestic Sci 5:225236.
Mikami N, Yoshimura J, Yamada H, Miyamoto J (1984a) Translocation and metabolism
of procymidone in cucumber and bean plants. J Pestic Sci 9:131136.

Photodegradation of Pesticides

177

Mikami N, Imanishi K, Yamada H, Miyamoto J (1984b) Photodegradation of the fungicide tolclofos-methyl in water and on soil surfaces. J Pestic Sci 9:215222.
Mikami N, Takahashi N, Yamada H, Miyamoto J (1985a) Separation and identification
of short-lived free radicals formed by photolysis of the pyrethroid insecticide fenvalerate. J Pestic Sci 16:101112.
Mikami N, Imanishi K, Yamada H, Miyamoto J (1985b) Photodegradation of fenitrothion in water and on soil surface, and its hydrolysis in water. J Pestic Sci 10:263
272.
Miller GC, Donaldson SG (1994) Factors affecting photolysis of organic compounds on
soils. In: Helz GR, Zepp RG, Crosby DG (eds) Aquatic and Surface Photochemistry.
Lewis, Boca Raton, FL, pp 97109.
Miller GC, Zepp RG (1979) Photoreactivity of aquatic pollutants sorbed on suspended
sediments. Environ Sci Technol 13:860863.
Miller GC, Zepp RG (1983) Extrapolating photolysis rates from the laboratory to the
environment. Residue Rev 85:89110.
Miller GC, Herbert VR, Miller WW (1989) Effect of sunlight on organic contaminants
at the atmosphere-soil interface. In: Reactions and Movement of Organic Chemicals
in Soils. Special Publication No. 622. Soil Science Society of America, Madison, WI,
pp 99110.
Miller LL, Nordblom GD, Yost GA (1974) Photochemistry of N-(-trichloromethyl-pmethoxybenzyl)-p-methoxyaniline. J Agric Food Chem 22:853855.
Minelli EV, Cabras P, Angioni A, Garau VL, Meils M, Pirisi FM, Cabitza F, Cubeddu
M (1996) Persistence and metabolism of fenthion in orange fruit. J Agric Food Chem
44:936939.
Misra B, Graebing PW, Chib JS (1997) Photodegradation of chloramben on a soil surface: a laboratory-controlled study. J Agric Food Chem 45:14641467.
Miyamoto J, Sato Y (1965) Determination of insecticide residue in animal and plant
tissues. Botyu-Kagaku 30:4549.
Moore WM, DuPont RR, McLean JE (1989) Soil phase photodegradation of toxic organics at contaminated disposal sites for soil renovation and groundwater quality protection. USGS/G-1304, No. PB89-237267. NTIS, Springfield, VA.
Mosier AR, Guenzi WD, Miller LL (1969) Photochemical decomposition of DDT by a
free-radical mechanism. Science 164:10831085.
Mostafa IY, Fakhr IMI, El-Zawahry YA (1974) Metabolism of organophosphorus insecticides. XV. Translocation and degradation of 32P-malathion in bean and cotton plants.
In: Proceedings, Comparative Studies on Food and Environmental Contamination.
IAEA, Vienna, pp 385392.
Moye HA, Winefordner JD (1965) Phosphorimetric study of some common pesticides.
J Agric Food Chem 13:516518.
Moye HA, Malagodi MH, Yoh J, Deyrup CL, Chang SM, Leibee GL, Ku CC, Wislocki
PG (1990) Avermectin B1a metabolism in celery: a residue study. J Agric Food Chem
38:290297.
Mueller TC, Moorman TB, Locke MA (1992) Detection of herbicides using fluorescence
spectroscopy. Weed Sci 40:270274.
Muller T, Maurer T, Kubiak R (1995) Metabolism and volatilization of parathion-methyl
under simulated outdoor conditions. Meded Fac Landbouww Univ Gent 60:541547.
Mumma RO, Khalifa S, Hamilton RH (1971) Spectroscopic identification of metabolites
of carbaryl in plants. J Agric Food Chem 19:445451.

178

T. Katagi

Murillo Pulgarin JA, Garca Bermejo LF (2002) Determination of the pesticide napropamide in soil, pepper, tomato by micelle-stabilized room-temperature phosphorescence. J Agric Food Chem 50:10021008.
Murthy NBK, Hustert K, Moza PN, Kettrup A (1998) Photodegradation of selected fungicides on soil. Fresenius Environ Bull 7:112117.
Nag SK, Dureja P (1996) Phototransformation of triadimefon on glass and soil surfaces.
Pestic Sci 48:247252.
Nag SK, Dureja P (1997) Photodegradation of azole fungicide triadomefon. J Agric Food
Chem 45:294298.
Nag-Chaudhuri J, Augenstein L (1964) Effect of physical environment on excited states
of amino acids and proteins. Targeted Diagnosis and Therapy (1964) 13:441452.
Nakajima A, Hidaka H (1993) Photosensitized oxidation of oleic acid, methyl oleate,
and olive oil using visible light. J Photochem Photobiol A Chem 74:189194.
Nambu K, Ohkawa H, Miyamoto J (1980) Metabolic fate of phenothrin in plants and
soils. J Pestic Sci 5:177197.
Nicholls CH, Leermakers PA (1971) Photochemical and spectroscopic properties of organic molecules in adsorbed or other perturbing polar environments. In: Pitts JN Jr,
Hammond GS, Noyes WA (eds) Advances in Photochemistry, vol 8. Wiley-Interscience, New York, pp 315336.
Nigg HN, Stamper JH, Knaak JB (1984) Leaf, fruit, and soil surface residues of carbosulfan and its metabolites in Florida citrus groves. J Agric Food Chem 32:8085.
Nilles GP, Zabik MJ (1974) Photochemistry of bioactive compounds. Multiphase photodegradation of basalin. J Agric Food Chem 22:684688.
Nilles GP, Zabik MJ (1975) Photochemistry of bioactive compounds. Multiphase photodegradation and mass spectral analysis of basagran. J Agric Food Chem 23:410415.
Nir S, Undabeytia T, Yaron-Marcovich D, El-Nahhal Y, Polubesova T, Serban C, Rytwo
G, Lagaly G, Rubin B (2000) Optimization of adsorption of hydrophobic herbicides
on montmorillonite preadsorbed by monovalent organic cations: interaction between
phenyl rings. Environ Sci Technol 34:12691274.
Niu J, Chen J, Henkelmann B, Quan X, Yang F, Kettrup A, Schramm KW (2003) Photodegradation of PCDD/Fs adsorbed on spruce (Picea abies (L.) Karst.) needles under
sunlight irradiation. Chemosphere 50:12171225.
Norris LA, Montgomery ML, Warren LE (1987) Triclopyr persistence in western Oregon
hill pastures. Bull Environ Contam Toxicol 39:134141.
Nutahara M, Murai T (1984) Accelerating effect of natural unsaturated fatty acids on
photodecomposition of chinomethionat (Morestan). J Pestic Sci 9:667674.
Ogawa K, Tsuda M, Yamaguchi F, Yamaguchi I, Misato T (1976) Metabolism of 2-secbutylphenyl N-methylcarbamate (Bassa, BPMC) in rice plants and its degradation in
soils. J Pestic Sci 1:219229.
Ohkawa H, Yoshihara R, Kohara T, Miyamoto J (1974a) Metabolism of m-tolyl Nmethylcarbamate (Tsumacide) in rats, houseflies and bean plants. Agric Biol Chem
38:10351044.
Ohkawa H, Mikami N, Miyamoto J (1974b) Photodecomposition of sumithion (O,Odimethyl-O-(3-methyl-4-nitrophenyl)phosphorothioate). Agric Biol Chem 38:2247
2255.
Ohkawa H, Nambu K, Miyamoto J (1980) Metabolic fate of fenvalerate (sumicidin) in
bean plants. J Pestic Sci 5:215223.
Ohsawa K, Casida JE (1979) Photochemistry of the potent knockdown pyrethroid kadethrin. J Agric Food Chem 27:11121120.

Photodegradation of Pesticides

179

Oliver BG, Cosgrove EG, Carey JH (1979) Effect of suspended sediments on the photolysis of organics in water. Environ Sci Technol 13:10751077.
Oltmans SJ (1981) Surface ozone measurements in clean air. J Geophys Res 86:1174
1180.
Oltmans SJ, Levy H II (1992) Seasonal cycle of surface ozone over the western North
Atlantic. Nature (Lond) 358:392394.
Ophoff FF, Stork A, Smelt J (1999) Volatilization of fenpropimorph under simulated
field conditions after application onto different plants. In: Human Environmental Exposure to Xenobiotics. Proceedings, 11th Symposium on Pesticide Chemistry. Goliardica Pavese, Pavia, Italy, pp 199209.
Osawa T (1994) Novel natural antioxidants for utilization in food and biological systems.
In: Uritani I, Garcia VV, Mendoza EMT (eds) Postharvest Biochemistry of Plant
Food: Materials in the Topics. Japan Science Society Press, Tokyo, pp 241251.
OToole JC, Cruz RT, Seiber JN (1979) Epicuticular wax and cuticular resistance in rice.
Physiol Plant 47:239244.
Paciolla MD, Davies G, Jansen SA (1999) Generation of hydroxyl radicals from metalloaded humic acids. Environ Sci Technol 33:18141818.
Parlar H (1980) Photochemistry at surfaces and interfaces. In: Hutzinger O (ed) The
Handbook of Environmental Chemistry. vol 2, part A. Springer-Verlag, Berlin, pp
145159.
Parlar H (1984) Geochemical induced degradation of environmental chemicals. Fresenius
Z Anal Chem 319:114118.
Parlar H (1990) The role of photolysis in the fate of pesticides. In: Hutson DH, Roberts
TR (eds) Progress in Pesticide Biochemistry and Toxicology, vol 7. Wiley, New
York, pp 245276.
Parlar H, Mansour M, Baumann R (1978) Photoreactions of hydroxychlordane in solution, as solids, and on surface of leaves. J Agric Food Chem 26:13211324.
Parochetti JV, Hein ER (1973) Volatility and photodecomposition of trifluralin, benefin
and nitralin. Weed Sci 21:469473.
Parochetti JV, Dec GW Jr (1978) Photodecomposition of eleven dinitroaniline herbicides. Weed Sci 26:153156.
Peacock GA, Riches MN, Wood S (1994) A new method for the evaluation of the photostability of crop protection compounds: the prediction of photostability in the field.
BCPC Monogr 59:251256.
Pelizzetti E, Carlin V, Maurino V, Minero C, Dolci M, Marchesini A (1990) Degradation
of atrazine in soil through induced photocatalytic processes. Soil Sci 150:523526.
Pelizzetti E, Minero C, Carlin V (1993) Photoinduced degradation of atrazine over different metal oxides. New J Chem 17:315319.
Pere E, Cardy H, Cairon O, Simon M, Lacombe S (2001) Quantitative assessment of
organic compounds adsorbed on silica gel by FT-IR and UV-VIS spectroscopies: the
contribution of diffuse reflectance spectroscopy. Vib Spectrosc 25:163175.
Petigara BR, Blough NV, Mignerey AC (2002) Mechanisms of hydrogen peroxide decomposition in soils. Environ Sci Technol 36:639645.
Piccinini P, Pichat P, Guillard C (1998) Phototransformations of solid pentachlorophenol.
J Photochem Photobiol A Chem 119:137142.
Pirisi FM, Cabras P, Garau VL, Melis M, Secchi E (1996) Photodegradation of pesticides. Photolysis rates and half-life of pirimicarb and its metabolites in reactions in
water and in solid phase. J Agric Food Chem 44:24172422.

180

T. Katagi

Pirisi FM, Angioni A, Cabizza M, Cabras P, Maccioni E (1998) Influence of epicuticular


waxes on the photolysis of pirimicarb in the solid phase. J Agric Food Chem 46:
762765.
Pirisi FM, Angioni A, Cabizza M, Cabras P, Cao CF (2001) Photolysis of pesticides:
influence of epicuticular waxes from Persica laevis DC on the photodegradation in
the solid phase of aminocarb, methiocarb and fenthion. Pestic Manag Sci 57:522526.
Pohlman AA, Mill T (1983) Peroxy radical interaction with soil constituents. Soil Sci
Soc Am J 47:922927.
Popendorf WJ, Leffingwell JT (1978) Natural variations in the decay and oxidation of
parathion foliar residues. J Agric Food Chem 26:437441.
Power JF, Sharma DK, Langford CH, Bonneau R, Joussot-Dubien J (1987) Laser flash
photolysis studies of a well-characterized soil humic substances. In: Zika RG, Cooper
WJ (eds) Photochemistry of Environmental Aquatic Systems. ACS Symposium Series
327. American Chemical Society, Washington, DC, pp 157173.
Prinn RG (1994) The interactive atmosphere: Global atmospheric-biospheric chemistry.
Ambio 23:5061.
Que Hee SS, Paine SH, Sutherland RG (1979) Photodecomposition of a formulated
mixed butyl ester of 2,4-dichlorophenoxyacetic acid in aqueous and hexane solutions.
J Agric Food Chem 27:7982.
Quinstad GB, Staiger LE (1984) Photodegradation of fluvalinate. J Agric Food Chem
32:11341138.
Quinstad GB, Staiger LE, Schooley DA (1975) Environmental degradation of the insect
growth regulator methoprene (isopropyl (2E,4E)-11-methoxy-3,7,11-trimethyl-2,4dodecadienoate). III. Photodecomposition. J Agric Food Chem 23:299303.
Racke KD (1993) Environmental fate of chlorpyrifos. Rev Environ Contam Toxicol 131:
1150.
Radler F, Horn DHS (1965) The composition of grape cuticle wax. Aust J Chem 18:
10591069.
Rajasekharan Pillai VN (1977) Role of singlet oxygen in the environmental degradation
of chlorthiamid to dichlobenil. Chemosphere 6:777782.
Rau H, Hormann M (1981) Kinetic resolution of optically active molecules and asymmetric chemistry: Asynmmetrically sensitized photolysis of trans-3,5-diphenylpyrazoline. J Photochem 16:231247.
Reichman R, Wallach R, Mahrer Y (2000a) A combined soil-atmospheric model for
evaluating the fate of surface-applied pesticides. 1. Model development and verification. Environ Sci Technol 34:13131320.
Reichman R, Mahren Y, Wallach R (2000b) A combined soil-atmospheric model for
evaluating the fate of surface-applied pesticides. 2. The effect of varying environmental conditions. Environ Sci Technol 34:13211330.
Reynolds G, Graham N, Perry R, Rice RG (1989) Aqueous ozonation of pesticides: A
review. Ozone Sci Eng 11:339382.
Rhodes RC (1977) Studies with manganese [14C]ethylenebis(dithiocarbamate) ([14C]maneb)
fungicide and [14C]ethylenethiourea ([14C]ETU) in plants, soil and water. J Agric Food
Chem 25:528533.
Richard C, Vialaton D, Aguer JP, Andreux F (1997) Transformation of monuron photosensitized by soil extracted humic substances: energy or hydrogen transfer mechanism? J Photochem Photobiol A Chem 111:265271.

Photodegradation of Pesticides

181

Riederer M, Schneider G (1990) The effect of the environment on the permeability and
composition of citrus leaf cuticles. II. Composition of soluble cuticular lipids and
correlation with transport properties. Planta (Berl) 180:154165.
Riskallah MR, Esaac EG, El-Sayed MM (1979) Photodegradation of leptophos. Bull
Environ Contam Toxicol 23:636641.
Riter RE, Adams VD, George DB, Kleine EA (1990) The effects of selected iron compounds on the sensitized photooxidation of bromacil. Chemosphere 21:717728.
Ritter WF, Johnson HP, Lovely WG (1973) Diffusion of atrazine, propachlor and diazinon in a silt loam soil. Weed Sci 21:381384.
Robberecht R, Caldwell MM (1980) Leaf ultraviolet optical properties along a latitudinal
gradient in the Arctic-alpine life zone. Ecology 61:612619.
Rodriguez E, Barrio RJ, Goicolea A, Peche R, de Balugera ZG, Sampedro C (2001)
Persistence of the insecticide Dimilin 45 ODC on conifer forest foliage in an Atlanticclimate ecosystem. Environ Sci Technol 35:38043808.
Rogers MAJ (1987) Singlet oxygen quantum yields. In: Heitz JR, Downum KR (eds)
Light-Activated Pesticides. ACS Symposium Series 339. American Chemical Society,
Washington, DC, pp 7697.
Romero E, Dios G, Mingorance MD, Matallo MB, Pena A, Sanchez-Rasero F (1998)
Photodegrdation of mecoprop and dichlorprop on dry, moist and amended soil surfaces exposed to sunlight. Chemosphere 37:577589.
Roof AAM (1982) Basic principles of environmental photochemistry. In: Hutzinger O
(ed) The Handbook of Environmental Chemistry, vol 2, part B. Springer-Verlag, Berlin, pp 117.
Rosen H, Margulies L (1991) Photostabilization of tetrahydro-2-(nitromethylene)-2H1,3-thiazine adsorbed on clays. J Agric Food Chem 39:13201325.
Rosen JD, Sutherland DJ (1967) The nature and toxicity of the photoconversion products
of aldrin. Bull Environ Contam Toxicol 2:19.
Ruggiero P (1999) Abiotic transformation of organic xenobiotics in soils: a compounding
factor in the assessment of bioavailability. In: NATO Science Series 2. Environmental
Security 64. Bioavailability of Organic Xenobiotics in the Environment. NATO,
Washington, DC, pp 159205.
Runeckles VC (1992) Uptake of ozone by vegetations. In: Lefohn AS (ed) Surface Level
Ozone Exposures and Their Effects on Vegetations. Lewis, Chelsea, MI, pp 157188.
Ruzo LO (1983) Involvement of oxygen in the photoreactions of cypermethrin and other
halogenated pyrethroids. J Agric Food Chem 31:11151117.
Ruzo LO, Casida JE (1979) Degradation of decamethrin on cotton plants. J Agric Food
Chem 27:572575.
Ruzo LO, Casida JE (1980) Pyrethroid photochemistry: mechanistic aspects in reactions
of the (dihalogenovinyl)cyclopropanecarboxylate substituent. J Chem Soc Perkin
Trans I 728732.
Ruzo LO, Casida JE (1981) Pyrethroid photochemistry: (S)--cyano-3-phenoxybenzyl
cis-(1R,3R,1R or S)-3-(1,2-dibromo-2,2-dihaloethyl)-2,2-dimethylcyclopropanecarboxylates. J Agric Food Chem 29:702706.
Ruzo LO, Casida JE (1982) Pyrethroid photochemistry: intramolecular sensitization and
photoreactivity of 3-phenoxybenzyl, 3-phenoxybenzoyl, and 3-benzoylbenzyl esters.
J Agric Food Chem 30:963966.
Ruzo LO, Casida JE (1985) Photochemistry of thiocarbamate herbicides: oxidative and
free radical processes of thiobencarb and diallate. J Agric Food Chem 33:272276.

182

T. Katagi

Ruzo LO, Zabik MJ, Schuetz RD (1974) Photochemistry of bioactive compounds: 1-(4chlorophenyl)-3-(2,6-dihalobenzoyl)ureas. J Agric Food Chem 22:11061108.
Ruzo LO, Holmstead RL, Casida JE (1977) Pyrethroid photochemistry: decamethrin.
J Agric Food Chem 25:13851394.
Ruzo LO, Gaughan LC, Casida JE (1980) Pyrethroid phorochemistry: S-bioallethrin.
J Agric Food Chem 28:246249.
Ruzo LO, Smith IH, Casida JE (1982) Pyrethroid photochemistry: photooxidation reactions of the chrysanthemates phenothrin and tetramethrin. J Agric Food Chem 30:
110115.
Ruzo LO, Krishnamurthy VV, Casida JE, Gohre K (1987) Pyrethroid photochemistry:
influence of the chloro(trifluoromethyl)vinyl substituent in cyhalothrin. J Agric Food
Chem 35:879883.
Sadeghi AM, Kissel DE, Cabrera ML (1989) Estimating molecular diffusion coefficients
of urea in unsaturated soil. Soil Sci Soc Am J 53:1518.
Saha T, Sukul P (1997) Metlaxyl: its persistence and metabolism in soil. Toxicol Environ
Chem 58:251258.
Samsonov YN, Makarov VI (1996) Kinetics and photophysical mechanism of sunlight
photolysis of unstable resmethrin and phenothrin in aerosols and their films. Bull
Environ Contam Toxicol 56:903910.
Samsonov YN, Pokrovskii LM (2001) Sensitized photodecomposition of high disperse
pesticide chemicals exposed to sunlight and irradiation from halogen or mercury
lamp. Atmos Environ 35:21332141.
Sanlaville Y, Guittonneau S, Mansour M, Feicht EA, Meallier P, Lettrup A (1996) Photosensitized degradation of terbutylazine in water. Chemosphere 33:353362.
Santoro A, Scopa A, Bufo SA, Mansour M, Mountacer H (2000) Photodegradation of
the triazole fungicide hexaconazole. Bull Environ Contam Toxicol 64:475480.
Sato K, Kato Y, Maki S, Matano O, Goto S (1985a) Penetration, translocation and metabolism of fungicide guazatine in dwarf apple trees. J Pestic Sci 10:8190.
Sato K, Kato Y, Maki S, Matano O, Goto S (1985b) Photolysis of fungicide guazatine
on glass surfaces. J Pestic Sci 10:91100.
Sauer HH (1972) Fate of formothion on bean plants in the greenhouse. J Agric Food
Chem 20:578583.
Saunders DG, Bret BL (1997) Fate of spinosad in the environment. Down Earth 52:
1420.
Schafmeier A, Emmelin C, Guittonneau S, Meallier P (1998) Influence of humic substances on the phenmedipham photodegradation. Fresenius Environ Bull 7:232237.
Schneiders GE, Koeppe MK, Naidu MV, Horne P, Brown AM, Mucha CF (1993) Fate
of rimsulfuron in the environment. J Agric Food Chem 41:24042410.
Scholz K, Reinhard F (1999) Photolysis of imidacloprid (NTN 33893) on the leaf surface
of tomato plants. Pestic Sci 55:652654.
Schonherr J, Riederer M (1989) Foliar penetration and accumulation of organic chemicals in plant cuticles. Rev Environ Contam Toxicol 108:170.
Schreiber L, Schonherr J (1993) Mobilities of organic compounds in reconstituted cuticular wax of barley leaves: determination of diffusion coefficients. Pestic Sci 38:353
361.
Schroeder J (1997) S-215 Regional research project final report. Behavior and fate of
selected sulfonylurea and imidazolinone herbicides in the southern environment.
Southern Cooperative Bulletin No. 385, Arkansas Agricultural Experiment Station,
Fayetteville, AR.

Photodegradation of Pesticides

183

Schroeder Kvien J, Banks PA (1985) Soil surface degradation of norflurazon. Weed Sci
Soc Am 25:95 (abstract).
Schuler F, Schmid P, Schlatter C (1998) Photodegradation of polychlorinated dibenzo-pdioxins and dibenzofurans in cuticular waxes of laurel cherry (Prunus laurocerasus).
Chemosphere 36:2134.
Schultz DP, Harman PD (1978) Hydrolysis and photolysis of the lampricide 2,5dichloro-4-nitrosalicylanilide (Bayer 73). Invest Fish Control 85:15.
Schwack W (1987) Photoreduction of parathion ethyl. Toxicol Environ Chem 14:6372.
Schwack W (1988) Photoinduced addition of pesticides to biomolecules. 2. Model reactions of DDT and methoxychlor with methyl oleate. J Agric Food Chem 36:645648.
Schwack W (1990) Photo-induced addition of pesticides to biomolecules. III. Model
reactions of folpet with cyclohexene. Z Lebensm-Unters-Forsch 190:420424.
Schwack W, Floer-Muller H (1990) Fungicides and photochemistry. Photodehalogenation of captan. Chemosphere 21:905912.
Schwack W, Hartmann M (1994) Fungicides and photochemistry: photodegradation of
the azole fungicide penconazole. Z Lebensm-Unters-Forsch 198:1114.
Schwack W, Kopf G (1992) Photodegradation of the carbamate insecticide propoxur. Z
Lebensm-Unters-Forsch 195:250253.
Schwack W, Kopf G (1993) Photodegradation of the carbamate insecticide pirimicarb.
Z Lebensm-Unters-Forsch 197:264268.
Schwack W, Andlauer W, Armbruster W (1994) Photochemistry of parathion in the plant
cuticle environment: model reaction in the presence of 2-propanol and methyl 12hydroxystearate. Pestic Sci 40:279284.
Schwack W, Bourgeois B, Walker F (1995a) Fungicides and photochemistry: photodegradation of the dicarboximide fungicide iprodione. Chemosphere 31:29933000.
Schwack W, Bourgeois B, Walker F (1995b) Fungicides and photochemistry: photodegradation of the dicarboximide fungicide procymidone. Chemosphere 31:40334040.
Schwack W, Walker F, Bourgeois B (1995c) Fungicides and photochemistry: photodegradation of the dicarboximide fungicide vinclozolin. J Agric Food Chem 43:3088
3092.
Schynowski F, Schwack W (1996) Photochemistry of parathion on plant surfaces: relationship between photodecomposition and iodine number of the plant cuticle. Chemosphere 33:22552262.
Scott HD, Phillips RE (1972) Diffusion of selected herbicides in soil. Soil Sci Soc Am
Proc 36:714719.
Scott HD, Phillips RE (1973) Self-diffusion coefficients of selected herbicides in water
and estimates of their transmission factors in soil. Soil Sci Soc Am Proc 37:965967.
Scrano L, Bufo SA, DAuria M, Emmelin C (1999) Photochemical behavior of oxyfluorfen: a diphenyl-ether herbicide. J Photochem Photobiol A Chem 129:6570.
Segura-Carretero A, Cruces-Blanco C, Canabate-Diaz B, Fernandez-S&acute;anchez JF,
Fernandez-Gutierrez A (2000) Heavy-atom induced room-temperature phosphorescence: a straightforward methodology for the determination of organic compounds in
solution. Anal Chim Acta 417:1930.
Sen A (1987) Chemical composition and morphology of epicuticular waxes from leaves
of Solanum tuberosum. Z Naturforsch 42c:11531158.
Senesi N, Loffredo E (1997) Minimizing environmental damage originating from pesticide utilization: abiotic photochemical control and remedies. In: Rosen D, Tel-Or E,
Hadar Y, Chen Y (eds) Developments in Plant and Soil Sciences. Modern Agriculture
and Environment, vol. 71. Kluwer, London, pp 4773.

184

T. Katagi

Senesi N, Miano TM (1995) The role of abiotic interactions with humic substances on
the environmental impact of organic pollutants. In: Huang PM, Berthelin J, Bollag
JM, McGill WB (eds) Environmental Impacts of Soil Component Interactions: Natural and Anthropogenic Organics, vol 1. CRC Press, Boca Raton, FL, pp 311335.
Senesi N, Schnitzer M (1977) Effects of pH, reaction time, chemical reduction and irradiation on ESR spectra of fulvic acid. Soil Sci 123:224234.
Senesi N, Testini C (1984) Theoretical aspects and experimental evidence of the capacity
of humic substances to bind herbicides by charge-transfer mechanism. Chemosphere
13:461468.
Senesi N, Miano TM, Provenzano MR, Brunetti G (1989) Spectroscopic and compositional comparative characterization of I.H.S.S. reference and standard fulvic and humic acids of various origin. Sci Total Environ 81/82:143156.
Sharma BK, Gupta N (1994) Photodegradation of the organophosphorus insecticide
phorate. Toxicol Environ Chem 41:249254.
Sharma KK, Chibber SS (1997) Photolysis of diniconazole-M under sunlight. Pestic Sci
49:115118.
Sherman DM (1989) Crystal chemistry, electronic structures, and spectra of Fe sites in
clay minerals: application to photochemistry and electron transport. In: Coyne LM,
Blake DF, McKeever SWS (eds) Spectroscopic Characterization of Minerals and
Their Surfaces. ACS Symposium Series 415. American Chemical Society, Washington, DC, pp 284309.
Singh HB, Ludwig FL, Johnson WB (1978) Tropospheric ozone: concentrations and
variabilities in clean remote atmospheres. Atmos Environ 12:21852196.
Slade M, Casida JE (1970) Metabolic fate of 3,4,5- and 2,3,5-trimethylphenyl methylcarbamates, the major constituents in landrin insecticide. J Agric Food Chem 18:467
474.
Smith AE, Grove J (1969) Photochemical degradation of diquat in dilute aqueous solution and on silica gel. J Agric Food Chem 17:609613.
Smith AM, Mao J, Doane RA, Kovacs MF Jr (1995) Metabolic fate of [14C] acrolein
under aerobic and anaerobic aquatic conditions. J Agric Food Chem 43:24972503.
Soeda Y, Kosaka S, Noguchi T (1972) The fate of thiophanate-methyl fungicide and its
metabolites on plant leaves and glass plates. Agric Biol Chem 36:931936.
Soeda Y, Ishihara K, Iwataki I, Kamimura H (1979) Fate of a herbicide 14C-alloxydimsodium in sugar beet. J Pestic Sci 4:121128.
Sogliero G, Eastwood D, Gilbert J (1985) A concise feature set for the pattern recognition of low-temperature luminescence spectra of hazardous chemicals. ASTM Special
Technical Publication 863. Advances in Luminescence Spectroscopy. American Society for Testing and Materials, Philadelphia, PA, pp 95115.
Somich CJ, Kearney PC, Muldoon MT, Elsasser S (1988) Enhanced soil degradation
of alachlor by treatment with ultraviolet light and ozone. J Agric Food Chem 36:
13221326.
Spear RC, Lee YS, Leffingwell JT, Jenkins D (1978) Conversion of parathion to paraoxon in foliar residues: effects of dust level and ozone concentration. J Agric Food
Chem 26:434436.
Spencer WF, Adams JD, Shoup TD, Spear RC (1980) Conversion of parathion to paraoxon on soil dusts and clay minerals as affected by ozone and UV light. J Agric Food
Chem 28:366371.

Photodegradation of Pesticides

185

Stamper JH, Nigg HN, Allen JC (1979) Organophosphate insecticide disappearance from
leaf surfaces: an alternative to first-order kinetics. Environ Sci Technol 13:1402
1405.
Stevenson FJ (1976) Organic matter reactions involving pesticides in soil. In: Kaufman
DD, Still GG, Paulson GD, Bandal SK (eds) Bound and Conjugated Pesticide Residues. ACS Symposium Series 29. American Chemical Society, Washington, DC, pp
180207.
Stiasni M, Deckers W, Schmidt K, Simon H (1969) Translocation, penetration, and metabolism of O-(4-bromo-2,5-dichlorophenyl)-O,O-dimethyl phosphorothioate (bromophos) in tomato plants. J Agric Food Chem 17:10171020.
Strek HJ (1998) Fate of chlorsulfuron in the environment. 1. Laboratory evaluations.
Pestic Sci 53:2951.
Suflita JM, Loll MJ, Snipes WC, Bollag JM (1981) Electron spin resonance study of
free radicals generated by a soil extract. Soil Sci 131:145150.
Sukul P, Spiteller M (2001) Influence of biotic and abiotic factors on dissipating metalaxyl in soil. Chemosphere 45:941947.
Sumida S, Yoshihara R, Miyamoto J (1973) Degradation of 3-(3,5-dichlorophenyl)-5,5dimethyloxazolidine-2,4-dione by plants, soil and light. Agric Biol Chem 37:2781
2790.
Suzuki M, Yamamoto Y (1974) Photodieldrin residues in field soils. Bull Environ Contam Toxicol 12:275280.
Svenningsson M (1988) Epi- and intracuticular lipids and cuticular transpiration rates of
primary leaves of eight barley (Hordeum vulgare) cultivars. Physiol Plant 73:512
517.
Takade DY, Seo MS, Kao TS, Fukuto TR (1976) Alteration of O,O-dimethyl S-[(carboethoxy)benzyl] phosphorodithioate (phenthoate) in citrus, water and upon exposure to air and sunlight. Arch Environ Contam Toxicol 5:6386.
Takagi K, Shichi T (2000) Photophysics and photochemistry in clay minerals. Mol Supramol Photochem 5:31110.
Takahashi N, Mikami N, Matsuda T, Miyamoto J (1985a) Photodegradation of the pyretyhroid insecticide cypermethrin in water and on soil surface. J Pestic Sci 10:629
642.
Takahashi N, Mikami N, Yamada H, Miyamoto J (1985b) Photodegradation of the pyrethroid insecticide fenpropathrin in water, on soil and on plant foliage. Pestic Sci 16:
119131.
Takahashi N, Ito M, Mikami N, Matsuda T, Miyamoto J (1988) Identification of reactive
oxygen species generated by irradiation of aqueous humic acid solution. J Pestic Sci
13:429435.
Tanaka AK, Umetsu N, Fukuto TR (1985) Metabolism of benfuracarb in young cotton,
bean and corn plants. J Agric Food Chem 33:10491055.
Tanaka FS, Wien RG, Mansager ER (1979) Effect of nonionic surfactants on the photochemistry of 3-(4-chlorophenyl)-1,1-dimethylurea in aqueous solution. J Agric Food
Chem 27:774779.
Tanaka FS, Wien RG, Mansager ER (1981) Survey for surfactant effects on the photodegradation of herbicides in aqueous media. J Agric Food Chem 29:227230.
Tanaka FS, Wien RG, Hoffer BL (1986) Photosensitized degradation of a homogeneous
nonionic surfactant: hexaethoxylated 2,6,8-trimethyl-4-nonanol. J Agric Food Chem
34:547551.

186

T. Katagi

Tanaka FS, Wien RG, Zaylskie RG (1991) Photolytic degradation of a homogeneous


Triton X nonionic syrfactant: nonaethoxylated p-(1,1,3,3-tetramethylbutyl)phenol.
J Agric Food Chem 39:20462052.
Theng YW, Chang IJ, Wang CM (1997) Direct evidence of clay-mediated charge transfer. J Phys Chem B 101:1038610389.
Thomas JK (1993) Physical aspects of photochemistry and radiation chemistry of molecules adsorbed on SiO2, -Al2O3, zeolites and clays. Chem Rev 93:301320.
Thomas SM, Harrison SK (1990) Surfactant-altered rates of chlorimuron and metsulfuron photolysis in sunlight. Weed Sci 38:602606.
Thomas-Smith TE, Blough NV (2001) Photoproduction of hydrated electron from constituents of natural water. Environ Sci Technol 35:27212726.
Tsao R, Eto M (1989) Chemical and photochemical transformation of the insecticide
cartap hydrochloride into nereistoxin. J Pestic Sci 14:4751.
Tsao R, Eto M (1990a) Photoreactions of the herbicide naproanilide and the effect of
some photosensitizers. J Environ Sci Health B 25:569585.
Tsao R, Eto M (1990b) Photolytic and chemical oxidation reactions of the insecticide
etofenprox. J Pestic Sci 15:405411.
Tsao R, Eto M (1991) Photolysis of flutolanil fungicide and the effect of some photosensitizers. Agric Biol Chem 55:763768.
Tsao R, Eto M (1994) Effect of some natural photosensitizers on photolysis of some
pesticides. In: Helz GR, Zepp RG, Crosby DG (eds) Aquatic and Surface Photochemistry. Lewis, Boca Raton, FL, pp 163171.
Tsao R, Hirashima A, Eto M (1989) Photolysis of the insecticide pyridafenthion and the
effect of some photosensitizers. J Pestic Sci 14:315319.
Turner DW (1959) Spectrophotometry in the far-ultraviolet region. Part II. Absorption
spectra of steroids and triterpenoids. J Chem Soc 3033.
Turner NC, Waggoner PE, Rich S (1974) Removal of ozone from the atmosphere by
soil and vegetation. Nature (Lond) 250:486489.
Turro NL (1978) Modern Molecular Photochemistry. Benjamin/Cummings, Menlo Park,
CA.
Uchida M, Ogawa K, Sugimoto T, Aizawa H (1983) Metabolism of flutolanil in rice
plant and cucumber. J Pestic Sci 8:537544.
Ueda K, Gaughan LC, Casida JE (1974) Photodecomposition of resmethrin and related
pyrethroids. J Agric Food Chem 22:212220.
Unai T, Tomizawa C (1986) Photodegradation of fenothiocarb on silica gel plate exposed
to sunlight. J Pestic Sci 11:363367.
Unai T, Tamaru H, Tomizawa C (1986) Translocation and metabolism of the acaricide
fenothiocarb in citrus. J Pestic Sci 11:347356.
Undabeytia T, Nir S, Tel-Or E, Rubin B (2000) Photostbilization of the herbicide norflurazon by using organoclays. J Agric Food Chem 48:47744779.
Van Koot IRY, Dijkhuizen T (1968) Light-transmission of dirty glass and clearing methods. Acta Hortic (ISHS) 6:97108.
Van Noort P, Lammers R, Verboom H, Wondergem E (1988) Rates of triplet humic acid
sensitized photolysis of hydrophobic compounds. Chemosphere 17:3538.
Vannelli JJ, Schulman EM (1984) Solid surface room-temperature phosphorescence of
pesticides. Anal Chem 56:10301033.
Vaugham PP, Blough NV (1998) Photochemical formation of hydroxyl radical by constituents of natural waters. Environ Sci Technol 32:29472953.

Photodegradation of Pesticides

187

Venkatesh R, Harrison SK (1999) Photolytic degradation of 2,4-D on Zea mays leaves.


Weed Sci 47:262269.
Verma NK, Pitliya RL, Vaidya VK, Ameta SC (1991) Photo-sensitized oxidation of
O,O-dimethyl O-4-nitro-m-tolyl phosphorothioate by singlet oxygen. Asian J Chem
3:260263.
Villemure G, Detellier C, Szabo AG (1986) Fluorescence of clay-intercalated methylviologen. J Am Chem Soc 108:46584659.
Voelker BM, Morel FMM, Sulzberger B (1997) Iron redox cycling in surface waters:
effects of humic substances and light. Environ Sci Technol 31:10041011.
Vogelmann TC, Bjorn LO (1984) Measurement of light gradients and spectral regime in
plant tissue with a fiber optic probe. Physiol Plant 60:361368.
Walia S, Dureja P, Mukerjee SK (1988) New photodegradation products of chlorpyrifos
and their detection on glass, soil, and leaf surfaces. Arch Environ Contam Toxicol
17:183188.
Walia S, Dureja P, Mukerjee SK (1989a) Photochemical transformation of phosalone.
Pestic Sci 25:19.
Walia S, Dureja P, Mukerjee SK (1989b) Photodegradation of the organophosphorus
insecticide iodofenphos. Pestic Sci 26:19.
Walker A, Crawford DV (1970) Diffusion coefficients for two triazine herbicides in six
soils. Weed Res 10:126132.
Walton TJ (1990) Waxes, cutin and suberin. Methods Plant Biochem 4:105158.
Watkins DAM (1974) Some implications of the photochemical decomposition of pesticides. Chem Ind 185190.
Watkins DAK (1987) Effects of leaf surfaces and plant waxes on rates of photodegradation of fenarimol. Asp Appl Biol 14:335346.
Weizmann A, Mazur Y (1958) Steroids and triterpenoids of citrus fruit. II. Isolation of
citrostadienol. J Org Chem 23:832834.
Wendlandt WWM, Hecht HG (1966) Reflectance Spectroscopy. Wiley-Interscience,
New York.
Wheeler OH, Mateos JL (1956) The ultraviolet absorption of isolated double bonds.
J Org Chem 21:11101112.
Whitaker BD, Schmidt WF, Kirk MC, Barnes S (2001) Novel fatty acid esters of pcoumaryl alcohol in epicuticular wax of apple fruit. J Agric Food Chem 49:3787
3792.
Willis GH, McDowell LL (1987) Pesticide persistence on foliage. Rev Environ Contam
Toxicol 100:2373.
Wilkinson F, Brummer JG (1981) Rate constants for the decay and reactions of the
lowest electronically excited singlet state of molecular oxygen in solution. J Phys
Chem Ref Data 10:809999.
Wolfe CJM, Halmans MTH, van der Heijde HB (1981) The formation of singlet oxygen
in surface waters. Chemosphere 10:5962.
Wolfe NL, Mingelgrin U, Miller GC (1990) Abiotic transformations in water, sediments,
and soils. In: Cheng HH (ed) Pesticides in the Soil Environment: Processes, Impact,
and Modeling. SSSA Book Series 2. Soil Science Society of America, Madison, WI,
pp 103168.
Wright WL, Warren GF (1965) Photochemical decomposition of trifluralin. Weeds 13:
329331.

188

T. Katagi

Wrzesinski CL, Arison BH, Smith J, Zinh DL, Vanden Heuvel WJA, Crouch LS (1996)
Isolation and identification of residues of 4-(epi-methylamino-4-deoxyavermectin
B1a benzoate from the surface of cabbage. J Agric Food Chem 44:304312.
Wuhrmann-Meyer K, Wuhrmann-Meyer M (1941) The absorption of ultraviolet light by
cuticular and wax layers of leaves. Planta (Berl) 32:4350.
Yamaoka K, Tsujino Y, Ando M, Nakagawa M, Ishida M (1988) Photolysis of DTP, the
herbicidal entity of pyrazolate, in water and on soil surface. J Pestic Sci 13:2937.
Yamazaki M, Sakai M, Goto F (1982) Behavior of acephate in tabacco plants treated as
wettable powder. J Pestic Sci 7:167173.
Yang X, Wang X, Kong L, Wang L (1999) Photolysis of chlorsulfuron and metsulfuronmethyl in methanol. Pestic Sci 55:75754.
Yih RY, Swithenbank C (1971) Identification of metabolites of N-(1,1-dimethylpropynyl)-3,5-dichlorobenzamide in soil and alfalfa. J Agric Food Chem 19:314319.
Yokley RA, Garrison AA, Wehry EL, Mamantov G (1986) Photochemical transformation of pyrene and benzo[a]pyrene vapor-deposited on eight coal stack ashes. Environ
Sci Technol 20:8690.
Yumita T, Yamamoto I (1982) Photodegradation of mepronil. J Pestic Sci 7:125131.
Yumita T, Shimazaki I, Miyamoto T, Yamamoto I (1984) Production of benazamide
and isoindoline type compounds on photodegradation of benzanilides. J Pestic Sci 9:
419423.
Zayed SMAD, Farghaly M, Hassan A (1978) Chemistry and toxicology of pesticide
chemicals. VII. Photodecompostion of leptophos. Isotopenpraxis 2:6870.
Zayed SMAD, Mostafa IY, El-Arab AE (1994) Degradation and fate of 14C-DDT and
14
C-DDE in Egyptian soil. J Environ Sci Health B 29:4756.
Zepp RG (1982) Experimental approaches to environmental photochemistry. In: Hutzinger O (ed) The Handbook of Environmental Chemistry, vol 2, part B. SpringerVerlag, Berlin, pp 1941.
Zepp RG (1988) Environmental photoprocesses involving natural organic matter. In:
Frimmel FH, Christman RF (eds) Humic Substances and Their Role in the Environment. Wiley, New York, pp 193214.
Zepp RG (1991) Photochemical conversion of solar energy in the environment. In: Pelizzetti E, Schiavello M (eds) Photochemical Conversion and Storage of Solar Energy.
Kluwer, Dordrecht, pp 497515.
Zepp RG, Cline DM (1977) Rates of direct photolysis in aquatic environment. J Agric
Food Chem 11:359366.
Zepp RG, Schlotzhauer PF (1981) Effects of equilibration time on photoreactivity of the
pollutant DDE sorbed on natural sediments. Chemosphere 10:453460.
Zepp RG, Wolfe NL, Baughman GL, Hollis RC (1977) Singlet oxygen in natural waters.
Nature (Lond) 267:421423.
Zepp RG, Baughman GL, Schlotzhauer PF (1981) Comparison of photochemical behavior of various humic substances in water: I. Sunlight induced reactions of aquatic
pollutants photosensitized by humic substances. Chemosphere 10:109117.
Zepp RG, Schlotzhauer PF, Sink RM (1985) Photosensitized transformation involving
electronic energy transfer in natural waters: role of humic substances. Environ Sci
Technol 19:7481.
Zepp RG, Braun AM, Hoigne J, Leenheer JA (1987) Photoproduction of hydrated electrons from natural organic solutes in aquatic environments. Environ Sci Technol 21:
485490.

Photodegradation of Pesticides

189

Zepp RG, Faust BC, Hoigne J (1992) Hydroxyl radical formation in aqueous reactions
(pH38) of iron (II) with hydrogen peroxide: the photo-Fenton reaction. Environ Sci
Technol 26:313319.
Zhang L, Brook JR, Vet R (2002) On ozone dry depositionwith emphasis on nonstomatal uptake and wet canopies. Atmos Environ 36:47874799.
Zhang M, Smyser BP, Shalaby LM, Boucher CR, Berg DS (1999) Lenacil degradation
in the environment and its metabolism in the sugar beets. J Agric Food Chem 47:
38433849.
Zongmao C, Haibin W (1997) Degradation of pesticides on plant surfaces and its prediction: a case study on tea plant. Environ Monit Assess 44:303313.
Manuscript received September 20, accepted October 20, 2003.

http://www.springer.com/978-0-387-20845-9

You might also like