Environmental Photochemistry of Altrenogest: Photoisomerization to a

Jun 29, 2016 - of this reaction were measured using transient absorption spectroscopy. Subsequent character- ization determined that this primary cycl...
0 downloads 11 Views 3MB Size
Subscriber access provided by La Trobe University Library

Article

Environmental Photochemistry of Altrenogest: Photoisomerization to a Bioactive Product with Increased Environmental Persistence via Reversible Photohydration Kristine H. Wammer, Kyler C. Anderson, Paul R. Erickson, Sarah Kliegman, Marianna E. Moffatt, Stephanie M. Berg, Jackie A. Heitzman, Nicholas Pflug, Kristopher McNeill, Dalma Martinovic-Weigelt, Ruben Abagyan, David M. Cwiertny, and Edward P. Kolodziej Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b02608 • Publication Date (Web): 29 Jun 2016 Downloaded from http://pubs.acs.org on July 8, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Environmental Science & Technology is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 29

Environmental Science & Technology

1

Environmental Photochemistry of Altrenogest:

2

Photoisomerization to a Bioactive Product with

3

Increased Environmental Persistence via Reversible

4

Photohydration

5

Kristine H. Wammer†,* Kyler C. Anderson,† Paul R. Erickson,‡ Sarah Kliegman,‡ • Marianna E.

6

Moffatt,† Stephanie M. Berg, † Jackie A. Heitzman,§ Nicholas C. Pflug,¥ Kristopher McNeill,‡

7

Dalma Martinovic-Weigelt,§ Ruben Abagyan,€ David M. Cwiertny,ǁ Edward P. Kolodziej,┴,#

8



9 10



Institute of Biogeochemistry and Pollutant Dynamics, ETH Zürich, CH-8092 Zürich, Switzerland

11

§

Department of Biology, University of St. Thomas, St. Paul, MN 55105, USA

12

¥

Department of Chemistry, University of Iowa, Iowa City, IA 52242, USA

13 14



15 16

ǁ

17 18



19 20

#

Department of Chemistry, University of St. Thomas, St. Paul, MN 55105, USA

Skaggs School of Pharmacy and Pharmaceutical Sciences, University of California, San Diego, 9500 Gilman, La Jolla, CA 92093-0747. Department of Civil and Environmental Engineering, University of Iowa, Iowa City, IA 52242, USA

Interdisciplinary Arts and Sciences, University of Washington, Tacoma, Tacoma, WA 98402 USA

Department of Civil and Environmental Engineering, University of Washington, Seattle, WA 98195-2700 USA

ACS Paragon Plus Environment

1

Environmental Science & Technology

Page 2 of 29

21

ABSTRACT: Despite its wide use as a veterinary pharmaceutical, environmental fate data is

22

lacking for altrenogest, a potent synthetic progestin. Here, it is reported that direct photolysis of

23

altrenogest under environmentally relevant conditions was extremely efficient and rapid (half-

24

life ~25 seconds). Photolysis rates (observed rate constant kobs = 2.7 ± 0.2 × 10-2 s-1) were

25

unaffected by changes in pH or temperature but were sensitive to oxygen concentrations (N2-

26

saturated kobs = 9.10 ± 0.32 × 10-2 s-1; O2-saturated kobs = 1.38 ± 0.11 × 10-2 s-1). The primary

27

photoproduct was identified as an isomer formed via an internal 2+2 cycloaddition reaction; the

28

triplet lifetime (8.4 ± 0.2 µs) and rate constant (8 × 104 s-1) of this reaction were measured using

29

transient absorption spectroscopy. Subsequent characterization determined that this primary

30

cycloaddition photoproduct undergoes photohydration. The resultant photostable secondary

31

photoproducts are subject to thermal dehydration in dark conditions, leading to reversion to the

32

primary cycloaddition photoproduct on a time scale of hours to days, with the photohydration

33

and dehydration repeatable over several light/dark cycles. This dehydration reaction occurs more

34

rapidly at higher temperatures and is also accelerated at both high and low pH values. In vitro

35

androgen receptor (AR)-dependent gene transcriptional activation cell assays and in silico

36

nuclear hormone receptor screening revealed that certain photoproducts retain significant

37

androgenic activity, which has implications for exposure risks associated with the presence and

38

cycling of altrenogest and its photoproducts in the environment.

39

ACS Paragon Plus Environment

2

Page 3 of 29

40 41

Environmental Science & Technology

INTRODUCTION The endocrine disrupting potential of synthetic progestins is of growing concern.1-3 For

42

example, exposure to levonorgestrel, a human contraceptive, at only 0.8 ng/L reduces fecundity

43

in fish.1 Because their reproductive effects are observed at sub-ng/L concentrations, a recent

44

review went so far as to state that synthetic progestins may be the pharmaceutical class of

45

greatest concern after ethinyl estradiol.3

46

Altrenogest (ALT, 17-α-allyl-17-β-hydroxyestra-4,9,11-trien-3-one), also known as

47

allyl-trenbolone or ally-trenbolone, is a potent synthetic progestin used as an equine and swine

48

veterinary pharmaceutical. Marketed commercially under various trade names including

49

Matrix® (for swine), Regu-Mate®, and Altresyn® (for horses), ALT is administered orally or in

50

animal feed to maintain pregnancy, synchronize estrus for breeding, or postpone estrus after

51

weaning.4-6 For estrous synchronization, swine receive 210-360 mg doses of ALT over 12-18

52

days, while in horses 230-400 mg is used over 15 days.7, 8 For pregnancy maintenance in horses,

53

an additional 3000-6000 mg of ALT may be dosed over 120 days, although in certain cases over

54

7000 mg total may be administered over the 300+ day gestation period.9

55

As with all veterinary pharmaceuticals, estimating annual mass loads of ALT is

56

complicated by the opaque nature of its production and usage. However, in the United States, it

57

is reasonable to estimate that manufacturer-recommended doses translate to several thousand kg

58

of annual ALT usage among ~66,000,000 swine and 3,600,000 horses (2012 data).10 An

59

advertisement for Regu-Mate® states that over 20 million doses were sold over its 30 years of

60

use.11 Although more recent comparable data does not seem to be available, ALT was easily the

61

most widely used veterinary steroid in the United Kingdom in 2000, with 380 kg applied to the

62

~10,000,000 animal swine herd.12 For comparison, the typical annual individual dose of ethinyl

ACS Paragon Plus Environment

3

Environmental Science & Technology

63

estradiol for contraceptive purposes in humans is under 8 mg, with approximately ~25 kg of

64

ethinyl estradiol used annually in the UK.12 Consistent with modern globalized industrial

65

agricultural practices, additional extensive use is expected in the EU and Asia. Nevertheless,

66

presumably because it is not a human pharmaceutical, ALT has been overlooked in recent

67

reviews of environmental progestin pharmaceuticals,3, 13

68

Page 4 of 29

Although available metabolism data for ALT is incomplete and fragmented,

69

environmental release through excretion via both urinary and fecal pathways is expected, where

70

at least some polar metabolites exhibit reduced progestin activity.14 In racehorses, no measurable

71

phase I metabolites (indicating structural transformation) were observed in urine, with only

72

phase II metabolism to glucuronide or sulfate conjugates detected.15 In swine, the short

73

environmental impact assessment statement for ALT indicates the excretion of several

74

conjugated Phase II metabolites, with at least one major metabolite forming active ALT upon

75

deconjugation and suggesting the strong possibility of glucuronide cleavage in environmental

76

systems.16

77

There is also scant data on ALT occurrence in the environment. One study detected

78

concentrations up to 1.1 ng/L at 4% of sites sampled in a survey of endocrine active chemicals in

79

Minnesota lakes. 17 There is no public domain literature that documents ALT’s environmental

80

fate, to the best of our knowledge. Exposure models used in environmental impact assessments

81

estimate concentrations below 20 ng/L in a static farm pond scenario,16 with 0.6-9 ng/L

82

concentrations possible in impacted receiving waters.18 As these values fall within concentration

83

ranges where other potent synthetic progestins impair fecundity in fish,3, 13 further investigation

84

of ALT’s environmental fate and impacts is merited.

ACS Paragon Plus Environment

4

Page 5 of 29

85

Environmental Science & Technology

Motivated by ALT’s structural similarity to trenbolone acetate (TBA, C18H22O2)

86

metabolites, in the present work we evaluated ALT’s photochemistry. ALT and TBA metabolites

87

share conjugated trienone substructures, and the photochemistry of the trienone moiety controls

88

the fate of TBA metabolites in shallow surface waters.19-21 Specifically, direct photolysis of TBA

89

metabolites proceeds via rapid photohydration, creating products that are not only photostable

90

but also capable of undergoing thermal dehydration under ambient conditions.22 These coupled

91

processes lead to photoproduct reversion that regenerates their parent structure and would be

92

expected to substantially increase exposure risks to TBA metabolites in affected receiving

93

waters.23

94

For ALT, we quantified phototransformation quantum yield and evaluated the impacts of

95

environmental factors (pH, temperature, and dissolved oxygen concentrations) on both

96

photolysis rate and photoproduct stability. Laser flash photolysis experiments coupled with NMR

97

spectroscopy and mass spectrometry were used to characterize major photoproducts and

98

determine the main phototransformation mechanism. Finally, to better understand the biological

99

implications of phototransformation products, an androgen receptor (AR)-dependent gene

100

transcriptional activation assay (MDA-kb2 cells) and an in silico nuclear hormone receptor

101

binding model were used to evaluate the bioactivity of ALT and its photoproducts.

102

EXPERIMENTAL METHODS

103

Photolysis Experiments and Photoproduct Characterization. Experiments used an

104

Atlas Suntest CPS+ solar simulator equipped with a xenon lamp and an Atlas UV Suntest filter,

105

to provide an emission spectrum (Figure S1) simulating that of natural sunlight at an irradiance

106

of 250 W/m2. Aqueous ALT solutions (10 µM, see supporting information [SI] for details about

107

all reagents and materials) were prepared in deionized water from methanol stock solutions (10

ACS Paragon Plus Environment

5

Environmental Science & Technology

Page 6 of 29

108

mM). Direct photolysis rates were measured as a function of solution pH (3.7 to 9.5), which was

109

adjusted using small amounts of HCl or NaOH, and temperature (30 to 44 °C). Solutions either

110

were sparged with nitrogen or oxygen or left open to air to test the effect of variable oxygen

111

concentration. To confirm environmental relevance, experiments were also conducted at a lower

112

concentration (~500 ng/L), in the presence of model organic matter (up to 80 mg/L of Fluka

113

Humic Acid), and with commercial preparations (Regu-Mate®, Matrix®, and Altresyn®), which

114

contain 0.22% ALT in an oil matrix (see SI for experimental details). Quantum yield

115

measurements and laser flash photolysis studies were also conducted as described in the SI.

116

Aqueous ALT concentrations were measured by high performance liquid chromatography

117

(HPLC), and NMR and mass spectrometry were used for photoproduct characterization; details

118

of analytical methods are found in the SI.

119

Photoproduct Stability Determination. The possibility of product-to-parent reversion

120

was examined by assessing photoproduct stability in the dark. ALT was photolyzed in deionized

121

water at 30 °C, after which the temperature (4-35 °C) and pH (2-12) of the photoproduct solution

122

was varied. The photoproduct mixture was stored in the dark, and aliquots were removed at

123

various time points for analysis of ALT and photoproduct concentrations.

124

Androgen Receptor (AR)-dependent gene transcriptional activation assay. To

125

evaluate product bioactivity, we quantified the androgenic activity of ALT and

126

ALT/photoproduct mixtures based on their ability to activate AR-mediated gene transcription in

127

MDA-kb2 cells stably transformed with the pMMTV neo luc reporter gene construct.24 Similar

128

assessment of progesterone receptor (PR)-mediated gene transcription would have been

129

desirable, but was beyond our capabilities. Thus, we focused on their androgenic activity because

130

certain progestins, especially early generation synthetic progestins, are also strong AR agonists

ACS Paragon Plus Environment

6

Page 7 of 29

Environmental Science & Technology

131

and this attribute also is implicated in their ability to disrupt endocrine function.25 Cell culturing

132

and transcriptional activation assays were performed following established procedures (described

133

in SI).24 Test solutions included dilution series of a testosterone (T) standard, ALT,

134

ALT/photoproduct mixtures, and appropriate control solutions that were reflective of

135

solvent/media ratios in test solutions. Two ALT/photoproduct mixtures were prepared from 10

136

µM ALT solutions photolyzed in the Suntest solar simulator as described previously, one with

137

significant ALT remaining (~12% of starting concentration, 40 seconds of photolysis) and one

138

photolyzed until no ALT remained (9 minutes of photolysis), in addition to a solution photolyzed

139

until no ALT remained and then allowed to thermally decompose in the dark for a minimum of

140

24 hours.

141

In Silico Modeling of ALT and ALT Photoproduct Binding with Nuclear Hormone

142

Receptors. To further explore conserved bioactivity in products, the potential binding of ALT

143

and ALT photoproducts to nuclear hormone receptors was assessed via virtual docking and

144

target screening.26-27 Full details are provided in the SI.

145 146

RESULTS AND DISCUSSION

147

Direct Photolysis Rate and Quantum Yield. ALT undergoes extremely rapid direct

148

photolysis under simulated sunlight, both at model and more environmentally relevant initial

149

concentrations (i.e., 10 µM and 1.6 nM, respectively; see Figure S2). For example, in deionized

150

water (pH 6) at 30 °C and open to air, photolysis occurred with a half-life of 27 seconds (kobs =

151

2.58 + 0.04 × 10-2 s-1). When corrected for light screening, the quantum yield (calculated as

152

described in the SI) was 0.38, suggesting ALT transformation is a very efficient photochemical

153

process. Because the direct photolysis occurs so rapidly, indirect photolysis is not expected to

ACS Paragon Plus Environment

7

Environmental Science & Technology

Page 8 of 29

154

significantly impact ALT fate in environmental systems. Further, such rapid and efficient

155

photolysis facilitates short half-lives even in highly turbid waters with elevated dissolved organic

156

matter concentrations (40 and 80 mg/L humic acid yielded half-lives of 6 and 10 min,

157

respectively; see Figure S3), as might be expected for systems impacted by agricultural runoff.

158

Solution pH (from pH ~4-9) and temperature (~30-44 °C) were found to have no

159

significant influence on photolysis rate constants, with half-lives of 23-28 seconds under all air-

160

saturated conditions tested. Values of kobs for photolysis under a range of pH, temperature, and

161

oxygen levels are given in Table S1, with data used to obtain kobs values from linear regression

162

analyses shown in Figures S4-S6. Notably, the presence or absence of oxygen in the system had

163

a significant effect on the ALT photolysis rate (Figure 1). At 30 °C, the reaction is half as fast

164

(t1/2 of 50 seconds, 1.38 ± 0.11 × 10-2 s-1) when purged with O2, and occurs almost too quickly to

165

quantify a rate coefficient (t1/2 of 8 seconds, kobs of 9.1 ± 0.3 × 10-2 s-1) when purged with N2.

166

We originally hypothesized that ALT’s photochemical behavior would be similar to TBA

167

metabolites (e.g., 17β-TBOH and 17α-trenbolone (17α-TBOH)) 19, which differ from ALT in

168

structure only at the C-17 position (see ALT structure in Figure 2 for atom numbering and ring

169

lettering conventions). However, some major differences were observed. ALT degradation

170

occurs far more rapidly than TBA metabolites, which typically undergo direct photolysis with

171

half-lives approximately two orders of magnitude longer. Photostability of the primary

172

photoproducts is also different. While TBA metabolites form photostable products (5- and 12-

173

hydroxylated trenbolone variants with λmax ~ 250 nm) because photohydration disrupts their

174

conjugated trienone system, ALT photolysis (Figure 1) yields one primary photoproduct that

175

remains photolabile. In addition, the strong dependence of ALT photolysis on oxygen

176

concentration is not observed for TBA metabolites, and further suggests a mechanism other than

ACS Paragon Plus Environment

8

Page 9 of 29

Environmental Science & Technology

177

a reversible photohydration for primary ALT photolysis. Of note is that photolysis of three

178

commercial ALT preparations (all in a vegetable oil matrix) resulted in the same dominant

179

primary photoproduct (Figure S7), albeit over significantly longer time scales (~2 orders of

180

magnitude).

181

Primary Photoproduct Characterization. Mass spectrometry revealed that the primary

182

photoproduct has identical mass to ALT ([M+H]+ = 311.2015), suggesting isomerization.

183

However, a significant shift in HPLC retention time (4.3 minutes for the primary photoproduct

184

versus 7.0 minutes for ALT on a reverse-phase C18 column) and major differences in the UV-

185

Vis absorbance spectra (Figure 3a) suggest that this isomerization must produce significant

186

structural changes. Indeed, the 1H NMR spectrum of the primary photoproduct is very different

187

from that of ALT. The vinyl region in ALT has three resonances characteristic of the endocyclic

188

olefins in its A and C rings (δ6.55, 6.40, 5.71 ppm) and three resonances from its terminal olefin

189

(δ5.98, 5.14, 5.10 ppm). In the photoproduct, only a single resonance remains (from the A ring

190

olefin, δ5.60 ppm), and the alkane region increases in complexity.

191

Collectively, these observations are consistent with ALT undergoing an intramolecular

192

photochemical [2+2] cycloaddition reaction between the pendant alkene (C17 substituent) and

193

the endocyclic alkene in the C ring of the steroid backbone (C11 – C12) to form new

194

cyclobutane and cyclopentane rings in the photoproduct [hereafter altrenogest cycloaddition

195

product (ALT-CAP), see Figure 2]. Although concerted [2+2] cycloaddition reactions are

196

thermally forbidden by the Woodward-Hoffmann rules, photochemical [2+2] cycloadditions are

197

allowed and have been observed in many systems since their discovery in 1908.28 Subsequent NMR analyses including 1H, 13C DEPT90 and 13C DEPT135, 1H-1H COSY,

198 199

1

H-13C HSQC, 1H-13C HMBC, and 1H-1H NOESY further support the formation of ALT-CAP

ACS Paragon Plus Environment

9

Environmental Science & Technology

Page 10 of 29

200

and were used to assign all of its 1H and 13C signals (see Table S5; Table S6 includes ALT

201

assignments for comparison). For example, the new sp3 CH signals at δ3.53, 2.80, 2.67, and CH2

202

signals at δ3.01, 1.55 ppm show connectivity and spatial relationships consistent with the

203

proposed formation of a cyclobutane ring via the photochemical [2+2] cycloaddition reaction.

204

The HSQC data, together with the 13C DEPT135, were used to establish the relationship between

205

the CH2 and CH carbons and their respective 1H signals, and the NOESY data suggest close

206

spatial relationships between the A, B, and E protons in the four-membered ring and the D

207

proton in the A-ring of the molecule.

208

The observed shift in the UV-Vis absorbance spectrum from a λmax of ~350 for ALT to a

209

λmax of ~320 for ALT-CAP (see Figure 3a) corroborates the NMR evidence for this structural

210

change. Woodward-Fieser rules are empirically-derived and predict λmax for certain classes of

211

compounds based on structure, including α,β-unsaturated ketones such as ALT and ALT-CAP.29-

212

30

213

and 321 for ALT-CAP). The key difference is an additional double bond extending conjugation

214

in the C ring for ALT compared to ALT-CAP, which according to the rules should result in a

215

decrease in λmax of 30 nm, as observed experimentally.

216

Both λmax values agree well with those predicted by the Woodward-Fieser rules (351 for ALT,

Laser Flash Photolysis. Fundamental insights explaining the differences in ALT (a 2+2

217

cycloaddition reaction) and 17β-TBOH (a photohydration reaction) phototransformation were

218

obtained with laser flash photolysis (LFP) to access their triplet-state absorbance and reactivity.

219

As shown in Figure 4a, following excitation ALT has a strong absorption around 430 nm with a

220

lifetime (τ) of 8.4 + 0.2 µs in N2-purged aqueous solution. The τ decreases dramatically in the

221

presence of O2, a known triplet quencher, indicating that the feature is a triplet-triplet absorption;

222

the kinetic trace of the decay is shown in Figure S8. Additionally, triplet energy transfer from

ACS Paragon Plus Environment

10

Page 11 of 29

Environmental Science & Technology

223

ALT to methylene blue, a low triplet energy acceptor, was observed when 100 µM of methylene

224

blue was present in solution.

225

Following excitation, 17β-TBOH has an equally strong absorption feature with a nearly

226

identical ∆A transient absorbance spectrum to ALT (Figures 4b and S8). This similarity is

227

explained by both compounds sharing the same trienone chromophore; the allylic group of ALT

228

does not substantially affect this chromophoric system (ALT and 17β-TBOH structures are both

229

shown on Figure 4). Despite these similarities, the observed τ of 17β-TBOH was 25.5 + 1 µs in

230

N2-purged solutions, a three-fold increase over ALT. The shorter lifetime for ALT results from

231

“triplet quenching” via the intramolecular cycloaddition reaction. Assuming that the τ of ALT

232

and 17β-TBOH would be the same in the absence of ALT’s C17 allylic group, we can use the

233

difference in lifetime to estimate the rate of the intramolecular reaction:

234

݇஺ = ்݇ + ݇ூ

(2)

235

where kA and kT are 1/τ for ALT and 17β-TBOH, respectively, and kI is the first-order rate of the

236

intramolecular reaction in ALT. Using equation (2) we calculated a rate constant of 8 x 104 s-1,

237

which falls in the same range of other similar photoinduced intramolecular reactions.31

238

Secondary Photoproduct Characterization and Effects of Solution Conditions. As

239

mentioned previously, ALT-CAP can be further photolyzed (t1/2 ~ 40 seconds, kobs ~ 2 × 10-2 s-1)

240

because its absorbance spectrum significantly overlaps the solar spectrum. One dominant

241

photoproduct is observed during ALT-CAP photolysis, and mass spectrometry reveals that this

242

ALT-CAP photoproduct (thus a secondary photoproduct of ALT) results from ALT-CAP

243

photohydration ([M+H]+ = 327.1963 Da, hereafter referred to as ALT-CAP-OH). Like

244

photoproducts of TBA metabolites, this photohydration is reversible; ALT-CAP-OH reverts to

ACS Paragon Plus Environment

11

Environmental Science & Technology

Page 12 of 29

245

ALT-CAP over ~48 hours in the dark (Figure 3b). This coupled photohydration-thermal

246

dehydration sequence can be repeated over multiple light-dark cycles, and the extent of reversion

247

appears independent of photoperiod. Although 3-4 minute photoperiods are shown in Figure 3b,

248

similar extents of reversion were observed for much longer (e.g., 8 h) photoperiods.

249

Temperature-dependent studies of the dark (thermal) stability of ALT-CAP-OH revealed

250

that the rate of ALT-CAP regrowth (i.e., reversion) increases with increasing temperature

251

(Figure 5a). The inset of Figure 5a shows Arrhenius plots for 17α-TBOH and ALT-CAP. It

252

illustrates that even though the absolute rate of ALT-CAP-OH dehydration is at least twice that

253

of the major 17α-TBOH photohydrate (i.e., 5-OH-17α-TBOH),21 the activation energies for

254

dehydration of both species are similar (~70 kJ/mol). We interpret this near identical effect of

255

temperature on 17α-TBOH and ALT-CAP reversion as evidence of structural similarity for their

256

photohydrates.

257

Generally, pH-dependent trends in ALT-CAP-OH dehydration also mirror those for TBA

258

metabolite photohydrates, although once again these rates are consistently faster for ALT-CAP-

259

OH. Dehydration is acid- and base-catalyzed, occurring faster at pH 5 and 9, respectively,

260

relative to neutral pH (Figure 5b). In fact, when ALT-CAP-OH solutions are adjusted to pH 2 or

261

12 immediately after their photoproduction, ALT-CAP regrowth is nearly instantaneous (Figure

262

S9).

263

This behavior closely resembles that previously reported for 17α-TBOH, which also

264

undergoes appreciable reversion at all pH values, whereas extensive reversion of 17β-TBOH

265

only occurs at acidic pH values (e.g., pH 5) (see inset of Figure 5b). As we have recently

266

detailed,22 the greater propensity for 17α-TBOH reversion can be attributed to its major

267

photohydrate forming via incorporation of H+ at C4 and OH- at C5. In contrast, the major

ACS Paragon Plus Environment

12

Page 13 of 29

Environmental Science & Technology

268

photohydrate of 17β-TBOH incorporates H+ at C4 and OH- at C12. Dehydration via water loss

269

across C4 and C5 occurs much more quickly relative to other possible photohydrate reactions

270

(e.g., subsequent hydration reactions) such that regeneration of 17α-TBOH can occur at all pH

271

values.

272

We interpret the shared pH- and temperature dependent trends for reversion of ALT-CAP

273

and 17α-TBOH as evidence that ALT-CAP photohydration also occurs via water addition across

274

the C4-C5 π-bond within its dienone moiety (see Figure 2). This is logical as the 2+2

275

cycloaddition eliminates the C11-C12 π-bond such that the alternate photohydration site

276

observed for 17β-TBOH is no longer available. Further, as with TBA metabolite

277

photohydrates,22 we presume that acid-catalyzed dehydration of ALT-CAP-OH occurs via a

278

unimolecular elimination reaction (E1), where loss of water yields a resonance-stabilized

279

carbocation intermediate that ultimately deprotonates to regenerate ALT-CAP. At higher pH,

280

dehydration most likely proceeds via enolate formation.

281

Androgenic Activity of ALT and ALT photoproducts. Androgenic activity of several

282

synthetic progestins has been widely documented.25, 32 For example, while such data is generally

283

limited for ALT, McRobb et al. used a yeast-based assay and determined that ALT was six times

284

more potent than testosterone (T) and its relative potency exceeded that of 17β-TBOH.33 In our

285

bioassays, ALT was androgenic at a level that was eight times less potent than T; EC50 values for

286

T and ALT were 0.37 and 2.9 nmol/L respectively (Figure S10, Table S3). 2-hydroxyflutamide

287

extinguished androgenic activity of ALT, indicating that observed activity was AR- and not GR-

288

mediated (data not shown). While estimates of androgenic potency were lower in the current

289

study than those reported by McRobb et al., both indicate that ALT is an effective androgen.

ACS Paragon Plus Environment

13

Environmental Science & Technology

290

Potency estimate discrepancies between yeast and mammalian cell assays are common and

291

expected; the causes of high variation between assays are addressed extensively elsewhere.34

Page 14 of 29

292

Androgenic activity was also detected in the photolyzed ALT samples after 40 seconds of

293

photolysis (peak of ALT-CAP formation, some ALT still present), 9 minutes of photolysis (ALT

294

and ALT-CAP completely transformed), and photolysis followed by 24 hours of thermal

295

decomposition (ALT-CAP regenerated, no ALT present); see Figure 6. We emphasize that only

296

the 40 second photolysis sample contained detectable ALT (see Table S2) indicating that

297

photoproducts, and not ALT, were responsible for induction of androgenic activity in the other

298

two samples.

299

Highly accurate estimation of relative potency of photoproducts in comparison to ALT or

300

the T standard is confounded by the change in concentrations over the course of the assay; ALT-

301

CAP-OH secondary photoproducts present at the beginning of the assay undergo thermal

302

dehydration and revert to ALT-CAP over the course of the assay (18 hours, 37°C). Furthermore,

303

complex matrices and chemical mixtures can hinder the ability to derive definitive, high-quality

304

estimates of androgenic potency in cell-based in vitro systems.35 Nevertheless, our data indicate

305

that ALT photoproducts exert significant androgenic activity. In fact, the product suite

306

collectively exhibits activity comparable to ALT. For example, if all of the activity observed in

307

the 9 minute photolysis sample is attributed to ALT-CAP, an EC50 value of 0.43 nmol/L is

308

calculated (Figure S10, Table S3). If this value were accurate, ALT-CAP would be a slightly less

309

potent androgen than T and slightly more potent than ALT. However, the observation that the

310

regeneration sample, which contains only ALT-CAP, results in a significantly lower response

311

(maximum response was too low for EC50 calculation, data not included on Figure S8), suggests

ACS Paragon Plus Environment

14

Page 15 of 29

Environmental Science & Technology

312

an overestimation of ALT-CAP activity due to experimental uncertainty and/or activity actually

313

attributable to secondary photoproducts.

314

As a final line of evidence for conserved bioactivity through ALT phototransformation,

315

we also employed in silico virtual target screening methods based upon three dimensional

316

receptor docking models. These in silico results are consistent with the aforementioned in vitro

317

bioassays, demonstrating that ALT is predicted to have a very high binding affinity (pKd = 8.5,

318

or Kd = 3.2 nM) for the androgen receptor (Table S4). This predicted AR affinity of ALT is on

319

par with its predicted (and expected) progesterone receptor binding (pKd = 9.4 with a confidence

320

interval of 1.2 pKd units). For ALT-CAP, outcomes of in silico modeling were also in agreement

321

with in vitro bioassay results. Specifically, virtual ligand screening data predict increased

322

androgen receptor affinity (pKd = 9.4) for ALT-CAP relative to ALT. Moreover, ALT-CAP also

323

is predicted to exhibit substantial retained affinity (pKd = 7.8) for the progesterone receptor,

324

although this outcome requires validation with in vitro assays targeting this receptor. Based on

325

the estimated confidence intervals of around one pKd unit for these predictions (corresponding to

326

an order of magnitude difference in the binding constant), the differences in pKd values around or

327

over 1 (e.g., 7.8 versus 9.4) are statistically significant.

328

Environmental Implications. These data demonstrate that assessment of risks associated

329

with ALT occurrence in sunlit aquatic environments needs to include ALT-CAP and the

330

secondary photoproducts. Across all tested conditions, including pH (3.7-9.5), temperature (30

331

to 44 °C), environmentally relevant (nanomolar) ALT concentrations, and high levels of natural

332

organic matter (up to 80 mg/L), ALT direct photolysis was efficient and rapid. Thus, we argue it

333

is not surprising that ALT is rarely observed in the environment. Given its very rapid photolysis

334

rate, ALT should be expected to degrade quickly and not be at all persistent, potentially even

ACS Paragon Plus Environment

15

Environmental Science & Technology

Page 16 of 29

335

over the time scales associated with drug handling and administration (notably, commercial

336

formulations of ALT provide no warning on their labels about its photochemical sensitivity). The

337

ALT photoproducts exhibit very efficient light/dark cycling, and at least some of these

338

transformation products exhibit biological activity at picomolar to nanomolar concentrations,

339

which is within the range anticipated to be present in agriculturally-impacted receiving waters.18

340

The nearly complete conservation of ALT-CAP mass through repeated photohydration-

341

dehydration cycles is notable. This suggests that ALT-CAP-OH, unlike TBA metabolite

342

photohydrates, is otherwise stable across all pH values and does not undergo subsequent

343

reactions to higher order hydrated products. Thus, other fate pathways (e.g., indirect photolysis,

344

biodegradation, sorption) will ultimately be required for attenuation of ALT-CAP mass from

345

ALT-impacted receiving waters.

346

Another noteworthy finding was the high androgenic potency of ALT and clear

347

androgenic activity of its more persistent photoproducts. These observations, along with the

348

likely co-occurrence of ALT and ALT photoproducts with other natural and synthetic androgens

349

in waters impacted by animal agriculture, suggest a more complex and nuanced interpretation of

350

environmental occurrence and associated ecological hazards for these agricultural

351

pharmaceuticals is necessary. Accordingly, future studies are needed to look for the presence of

352

these transformation products of potential concern, especially those such as ALT-CAP, in

353

addition to studies that will build upon these aqueous solution studies to understand more fully

354

what will happen to these molecules in more complex matrices. These represent clear examples

355

of what we have termed “environmental designer drugs”35, transformation products arising from

356

environmental processes that retain substantial structure and conserved bioactivity, yet are

ACS Paragon Plus Environment

16

Page 17 of 29

Environmental Science & Technology

357

otherwise invisible to standard, highly selective analytical methodologies focused on detection of

358

parent compounds.

359

ASSOCIATED CONTENT

360

Supporting Information. The Supporting Information is available free of charge on the ACS

361

Publications website at DOI: (TBD).

362

Supplementary text (Materials, Altrenogest Photolysis at Environmentally Relevant

363

Concentrations, Altrenogest Photolysis with Humic Acid, Photolysis of Commercial Oil

364

Solutions, Quantum Yield Measurements, Laser Flash Photolysis, Analytical Methods,

365

AR Assay Methods, In Silico Modeling of ALT and ALT Photoproduct Binding with

366

Nuclear Hormone Receptors, Quantum Yield Calculation); solar simulator spectral output

367

(Figure S1); ALT photolysis at low concentration and in the presence of humic acid

368

(Figures S2 and S3), direct photolysis rate constants and data used to obtain rate

369

constants (Table S1, Figures S4, S5, and S6); example chromatograms in water and oil

370

matrices (Figure S7), kinetic traces of triplet state decay (Figure S8), ALT-CAP

371

regeneration at pH 2 and 12 (Figure S9), ALT and ALT-CAP concentrations for AR

372

assay (Table S2), dose-response curves for AR assay (Figure S10), calculated EC50

373

values (Table S3), results of in silico virtual docking and target screening (Table S4), and

374

NMR shifts and assignments (Tables S5, S6).

375

AUTHOR INFORMATION

376

Corresponding Author

377

*Phone: +1-651-962-5574; fax +1-651-962-5201; email: [email protected]

ACS Paragon Plus Environment

17

Environmental Science & Technology

Page 18 of 29

378

Present Addresses

379



380

Notes

381

The authors declare no competing financial interest.

382

ACKNOWLEDGMENTS

383

The authors would like to thank Mr. Nicholas Cipoletti, Ms. Abigail Lukowicz, and Dr. Rachel

384

Lundeen for assistance with data collection, Dr. Dwight Stoll with assistance in method

385

development for oil extraction, and Dr. Kirk Heisterkamp, DVM, for assistance with acquisition

386

of Regu-Mate® and Altresyn®. Funding support was provided by the National Science

387

Foundation (CBET- 1335711 and CBET- 1336165) and the University of St. Thomas Grants and

388

Research Office.

389

REFERENCES

Current address for SK: Department of Chemistry, Reed College, Portland, OR 97202

390 391 392

(1) Zeilinger, J.; Steger-Hartmann, T.; Maser, E.; Goller, S.; Vonk, R.; Länge, R. Effects of synthetic gestagens on fish reproduction. Environ. Toxicol. Chem. 2009, 28(12), 2663−2670.

393 394 395

(2) Säfholm, M.; Ribbenstedt, A.; Fick, J.; Berg, C. Risks of hormonally active pharmaceuticals to amphibians: a growing concern regarding progestagens. Philos. T. Roy. Soc. A. 2014, 369, 20130577.

396 397 398

(3) Kumar, V.; Johnson, A.C.; Trubiroha, A.; Tumova, J.; Ihara, M.; Grabic, R.; Kloas, W.; Tanaka, H.; Kroupova, H.K. The challenge presented by progestins in ecotoxicological research: A critical review. Envrion. Sci. Techol. 2015, 49, 2625-2638.

399 400 401

(4) Squires, E.L.; Heesemann, C.P.; Webel, S.K.; Shideler, R.K.; Voss, J.L. Relationship of altrenogest to ovarian activity, hormone concentrations and fertility of mares. J. Anim. Sci. 1983, 56, 901-910.

402 403 404

(5) van Leeuwen, J.J.J.; Williams, S.I.; Martens, M.R.T.M.; Jourquin, J.; Driancourt, M.A.; Kemp, B.; Soede, N.M. The effect of postweaning altrenogest treatments of primiparous sows on follicular development, pregnancy rates, and litter sizes. J. Anim. Sci. 2011, 89, 397-403.

ACS Paragon Plus Environment

18

Page 19 of 29

Environmental Science & Technology

405 406 407

(6) Willmann, C.; Schuler, G.; Hoffmann, B.; Parvizi, N.; Aurich, C. Effects of age and altrenogest treatment on conceptus development and secretion of LH, prosgesterone, and eCG in early-pregnant mares. Theriogenology 2011, 75, 421-428.

408 409

(7) Regu-Mate® Product label; http://www.regu-mate.com/label.asp, accessed November 9, 2015.

410 411 412

(8) Estienne, M.J.; Harper, A.F.; Horsley, B.R.; Estienne, C.E.; Knight, J.W. Effects of P.G. 600 on the onset of estrus and ovulation rate in gilts treated with Regu-mate. J. Anim. Sci. 2001, 79, 2757-2761.

413 414

(9) Squires, E.L. Hormonal manipulation of the mare: A review. J. Equine Vet.Sci. 2008, 28, 627-634.

415

(10) United States Department of Agriculture. 2012 Census of Agriculture. AC-12-A-51, 2014.

416 417

(11) Merck advertisement, http://www.miskfortune.com/regu-mate/, accessed November 9, 2015.

418 419 420

(12) Boxall, A.B.A.; Fogg, L., Blackwell, P.A., Kay, P., Pemberton, E.J. Review of veterinary medicines in the environment. R&D Technical Report P6-012/8/TR. Environment Agency, Bristol, UK, 2002.

421 422 423

(13) Orlando, E.F.; Ellestad, L.E. Sources, concentrations, and exposure effects of environmental gestagens on fish and other aquatic wildlife, with an emphasis on reproduction. Gen. Comp. Endocr. 2014, 203, 241-249.

424 425

(14) European Medicines Agency. European public MRL assessment report (EPMAR): Altrenogest (equidae and porcine species). EMA/CVMP/487477/2011, London, UK, 2012.

426 427 428 429

(15) Lampinen-Salomonsson, M.L.; Beckman, E.; Bondesson, U.; Hedeland, M. Detection of altrenogest and its metabolites in post administration horse urine using liquid chromatography tandem mass-spectrometry – increased sensitivity by chemical derivatisation of the glucoronic acid conjugate. J. Chromatogr. B. 2006, 833, 245-256.

430 431

(16) Intervet, Inc. Finding of no significant impact for MATRIXTM (Altrenogest) solution 0.22% for cycling gilts. Millsboro, DE 2003

432 433 434

(17) Pharmaceuticals and Endocrine Active Chemicals in Minnesota Lakes; Document number tdr-g1-16; Minnesota Pollution Control Agency: Saint Paul, MN, 2013; http://www.pca.state.mn.us.

435 436 437

(18) European Medical Agency Environmental Impact Assessment, 2013; http://www.ema.europa.eu/docs/en_GB/document_library/Referrals_document/Suifertil_4_mg/ WC500153356.pdf, accessed November 9, 2015.

438 439 440

(19) Qu, S.; Kolodziej, E.P.; Cwiertny, D.M. Phototransformation rates and mechanisms for synthetic hormone growth promoters used in animal agriculture. Environ. Sci. Technol. 2012, 46, 13202-13211.

ACS Paragon Plus Environment

19

Environmental Science & Technology

Page 20 of 29

441 442 443 444

(20) Kolodziej, E.P.; Qu, S.; Forsgren, K.L.; Long, S.A.; Gloer, J.B.; Jones, G.D.; Schlenk, D.; Baltrusaitis, J.; Cwiertny, D.M. Identification and environmental implications of phototransformation products of trenbolone acetate metabolites. Environ. Sci. Technol. 2013, 47, 5031-5041.

445 446 447

(21) Qu, S.; Kolodziej, E. P.; Long, S. A.; Gloer, J. B.; Patterson, E. V.; Baltrusaitis, J.; Jones, G. D.; Benchetler, P. V.; Cole, E. A.; Kimbrough, K. C.; Tarnoff, D. M. Product-to-parent reversion of trenbolone: unrecognized risks for endocrine disruption. Science 2013, 342, 347-351.

448 449 450 451

(22) Baltrusaitis, J.; Patterson, E. V.; O'Connor, M.; Qu, S.; Kolodziej, E. P.; Cwiertny, D. M. Reversible photohydration of trenbolone acetate metabolites: Mechanistic understanding of product-to-parent reversion through complementary experimental and theoretical approaches. Environ. Sci. Technol. 2016, in press, DOI: 10.1021/acs.est.5b03905.

452 453 454

(23) Ward, A.S.; Cwiertny, D.M.; Kolodziej, E.P.; Brehm, C.C. Coupled reversion and streamhyporheic processes increase environmental persistence of trenbolone metabolites. Nature Communications. 2015, 6, 7067, doi:10.1038/ncomms8067.

455 456 457

(24) Wilson, V.S.; Bobseine, K.; Lambright, C.R.; Gray, Jr. L.E. A novel cell line, MDA-kb2, that stably expresses an androgen- and glucocorticoid-responsive reporter for the detection of hormone receptor agonists and antagonists. Toxicol. Sci. 2002, 66, 69-81.

458 459 460 461

(25) Ellestad, L.E.; Cardon, M.; Chambers, I.G.; Farmer, J.L.; Hartig, P.; Stevens, K.; Villeneuve, D.L.; Wilson, V.; Orlando, E.F. Environmental gestagens activate fathead minnow (Pimephales promelas) nuclear progesterone and androgen receptors in vitro. Environ. Sci. Technol. 2014, 48, 8179-87.

462 463

(26) Chen, Y.C.; Totrov, M.; Abagyan, R. Docking to multiple pockets or ligand fields for screening, activity prediction and scaffold hopping. Future Med. Chem. 2014, 6, 1741-1755.

464 465 466

(27) McRobb, R.M.; Kufareva, I.; Abagyan, R. In silico identification and pharmacological evaluation of novel endocrine disrupting chemicals that act via the ligand-binding domain of the estrogen receptor alpha. Toxicol. Sci. 2014, 141, 188-197.

467 468

(28) Wang, Z. Comprehensive Organic Name Reactions and Reagents; John Wiley and Sons: Hoboken, NJ, 2010.

469 470

(29) Woodward, R.B. Structure and the Absorption Spectra of α,β-Unsaturated Ketones. J. Am. Chem. Soc. 1941 63, 1123-1126.

471 472

(30) Fieser, L.F.; Fieser, M.; Rajagopalan, S. Absorption Spectrocsopy and the Structures of the Diosterols. J. Org. Chem. 1948 13, 800-806.

473 474

(31) Wagner, P.J.; Nahm, K. Regiospecific intramolecular reaction of an alkene group with the benzene ring of a triplet ketone. J. Am. Chem. Soc. 1987 109, 4404-4405.

475 476 477

(32) Moore, N.L.; Hickey, T.E.; Butler, L.M.; Tilley, W.D. Multiple nuclear receptor signaling pathways mediate the actions of synthetic progestins in target cells. Mol. Cell. Endocrinol. 2012 357, 60-70.

ACS Paragon Plus Environment

20

Page 21 of 29

Environmental Science & Technology

478 479 480

(33) McRobb, L., Handelsman, D.J., Kazlauskas, R.; Wilkinson, S.; McLeod, M.D.; Heather, A.K. Structure-activity relationships of synthetic progestins in a yeast-based in vitro androgen bioassay. J. Steroid Biochem. 2008 110, 39-47.

481 482 483

(34) Leusch, F.D.L.; De Jager, C.; Levi, Y.; Lim, R.; Puijker, L.; Sacher, F.; Tremblay, L.A.; Wilson, V.S.; Chapman, H.F. Comparison of five in vitro bioassays to measure estrogenic activity in environmental waters. Environ. Sci. Technol. 2010 44, 3853-3860.

484 485

(35) Cwiertny, D.M.; Snyder, S.A.; Schlenk, D.; Kolodziej, E.P. Environmental designer drugs: When transformation may not eliminate risk. Envrion. Sci. Techol. 2014, 48, 11737-11745.

486 487

Figure legends

488

Figure 1. Direct photolysis of 10 µM ALT at pH 6 and 30 °C in deionized water left open to the

489

air (a), purged with oxygen (b), and purged with nitrogen (c). Red circles are peak area of ALT

490

(at 350 nm); blue hollow squares are peak area of primary photoproduct (at 320 nm). Inset on

491

panel (c) illustrates a first order kinetic fit of the data; slopes are 9.1 + 0.32 x 10-2 s-1 for N2, 2.48

492

+ 0.04 x 10-2 s-1 for air, and 1.38 + 0.11 x 10-2 s-1 for O2.

493 494

Figure 2. Proposed pathway for photolysis of ALT. Photoisomerization occurs via a 2+2

495

cycloaddition to form the primary photoproduct (ALT-CAP), which subsequently undergoes a

496

reversible photohydration. Steroid ring lettering and atom numbering conventions are shown for

497

ALT.

498 499

Figure 3. (a) Absorbance spectra for a 10 µM ALT solution. A (blue) is an unphotolyzed

500

solution. B (green) has been photolyzed for 120 seconds and contains a mixture of primary

501

photoproduct and ALT. C (red) has gone through multiple photolysis/dark reversion cycles and

502

has only primary photoproduct. (b) Photolysis/dark reversion cycling of the same 10 µM ALT

503

solution. The spectra on panel (a) correspond to points A, B, and C. Black circles are ALT (peak

ACS Paragon Plus Environment

21

Environmental Science & Technology

504

areas at 354 nm, primary x-axis), red circles are primary photoproduct (peak areas at 320 nm,

505

secondary x-axis). Photolysis times are in seconds; reversion times are in hours.

Page 22 of 29

506 507

Figure 4. 2-D contour plots of transient absorption data after excitation for 100 µM solutions of

508

ALT (a) and 17β-TBOH (b). The color scale represents transient absorption (∆ Abs) intensity.

509 510

Figure 5. (a) Dependence on temperature of primary photoproduct (ALT-CAP) regeneration due

511

to thermal decomposition of secondary photoproducts (photohydrates). ALT-CAP peak area (at

512

320 nm) is shown as percent of initial ALT concentration (at 354 nm). Inset shows Arrhenius

513

plots for ALT-CAP and 17α−ΤΒΟΗ regeneration rates. (b) Dependence on pH of ALT-CAP

514

regeneration from thermal decomposition of secondary photoproducts. Inset shows analogous

515

data (red = pH 5, blue = pH 7, green = pH 9) for regeneration of 17α−TBOH and

516

17β−TBOH

517 518

Figure 6. Androgenic activity (expressed as a percent of maximum T response) of unphotolyzed

519

and photolyzed samples of a 10 µM ALT solution. Samples include ALT, a mixture of ALT and

520

its photoproducts (40s, 0.12 µM ALT remaining), photoproducts only after 9 min of photolysis,

521

and primary photoproduct only (solution photolyzed until no ALT remained and then allowed to

522

thermally decompose in the dark for a minimum of 24 hours.) All samples were serially diluted

523

(10-fold each time) in assay media.

ACS Paragon Plus Environment

22

Page 23 of 29

Environmental Science & Technology

a) Air

Peak Area (mAU)

b) O2

ln([ALT]t/[ALT]0)

c) O2

N2

Air N2 Time(sec)

Time

524 525

Figure 1.

ACS Paragon Plus Environment

23

Environmental Science & Technology

Page 24 of 29

526 527

Figure 2.

ACS Paragon Plus Environment

24

Page 25 of 29

Environmental Science & Technology

a)

Absorbance (AU)

A

C

B

Wavelength (nm)

A

Altrenogest PA (mAU)

B

Time (s)

528 529

C

Time (h) Time (s)

Time (h)

Time (s)

Photoproduct PA (mAU)

b)

Time (h)

Figure 3.

530

ACS Paragon Plus Environment

25

Environmental Science & Technology

Page 26 of 29

531 532

Figure 4.

ACS Paragon Plus Environment

26

Page 27 of 29

Environmental Science & Technology

ALT 17α -TB OH

ln(k)

Percent of Initial Altrenogest Peak Area

a)

45 °C

21 °C 1/T (1/K x 103)

4 °C k

pH 5

pH 7 pH 9

Percent of Initial Mass

Percent of Initial Altrenogest Peak Area

b)

17α α-TBOH 17β β-TBOH

Time

Time (h)

533 534

Figure 5.

ACS Paragon Plus Environment

27

Environmental Science & Technology

Page 28 of 29

535 536

% Maximum Testosterone Response

140

ALT + photoproducts (40 s)

120

Photoproducts (9 min)

100

ALT-CAP (regeneration) 80 60 40 20 0 1.E+01

1.E+02

1.E+03

1.E+04

1.E+05

1.E+06

1.E+07

1.E+08

Dilution Factor

537 538

ALT

Figure 6.

539

ACS Paragon Plus Environment

28

Page 29 of 29

Environmental Science & Technology

TOC Art

Altrenogest

Primary Photoproduct

Time (s)

Time (h)

Time (s)

Time (h)

Time (s)

ACS Paragon Plus Environment

Time (h)