+ All Categories
Home > Documents > Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

Date post: 11-Sep-2021
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
128
Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via direct and indirect photochemistry Author(s): Erickson, Paul R. Publication Date: 2014 Permanent Link: https://doi.org/10.3929/ethz-a-010195680 Rights / License: In Copyright - Non-Commercial Use Permitted This page was generated automatically upon download from the ETH Zurich Research Collection . For more information please consult the Terms of use . ETH Library
Transcript
Page 1: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

Research Collection

Doctoral Thesis

Transformation mechanisms of organic micropollutants viadirect and indirect photochemistry

Author(s): Erickson, Paul R.

Publication Date: 2014

Permanent Link: https://doi.org/10.3929/ethz-a-010195680

Rights / License: In Copyright - Non-Commercial Use Permitted

This page was generated automatically upon download from the ETH Zurich Research Collection. For moreinformation please consult the Terms of use.

ETH Library

Page 2: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 

DISS. ETH NO. 21854

Transformation mechanisms of organic micropollutants via

direct and indirect photochemistry

A dissertation submitted to

ETH ZURICH

For the degree of

Doctor of Sciences

Presented by

Paul Ragnar Erickson Jr.

MSc Chemistry, University of Minnesota-Twin Cities

born 12 March 1985

citizen of the

United States of America

Accepted on the recommendation of

Prof. Dr. Kristopher McNeill, examiner

Prof. Dr. David Blank, co-examiner

Dr. Silvio Canonica, co-examiner

2014

Page 3: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 

Page 4: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 

Table of Contents

Summary.......................................................................................................................1  Zusammenfassung .......................................................................................................5  Chapter 1- Introduction

Introduction......................................................................................................................10 Goals and thesis overview ...............................................................................................17

Chapter 2- Photochemical formation of brominated dioxins and other products of concern from hydroxylated polybrominated diphenyl ethers (OH-PBDEs)

Abstract ............................................................................................................................26 Introduction......................................................................................................................27 Experimental ....................................................................................................................30 Results and Discussion ...................................................................................................34 Appendix..........................................................................................................................46

Chapter 3- Disparate controlling factors in the rates of oxidation of anilines and phenols by triplet methylene blue in aqueous solution  

Abstract ............................................................................................................................68 Introduction......................................................................................................................69 Materials and Methods.....................................................................................................70 Results and Discussion ...................................................................................................74 Implications and Conclusions ..........................................................................................92 Appendix..........................................................................................................................98

Chapter 4- Investigating the mechanism of the photochemical chlorination of pyrene in aqueous solution  

Abstract ..........................................................................................................................104 Introduction....................................................................................................................105 Methods .........................................................................................................................106 Results and Discussion .................................................................................................107

Chapter 5- Conclusions and outlook......................................................................118  Acknowledgements ..................................................................................................123  Curriculum vitae.......................................................................................................124  

Page 5: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 

Page 6: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  1  

Summary

The release of organic pollutants into the environment is a global problem and

can have adverse effects on both human and environmental health. To accurately

access the health and environmental risks associated with the use and disposal of

pharmaceuticals, pesticides, and other industrially produced chemicals, it is important

to know their fate in the environment. To properly determine the risks of a given

compound in the environment, we must first determine which processes are important

for its fate and account for the products of its transformation, which may also be of

environmental concern. For many environmental pollutants one route of

transformation is exposure to sunlight or reactive intermediates that are generated

photochemically in the environment. Numerous possible photoreactions can affect the

fate of organic compounds. To achieve the goal of pollutant fate prediction, the

precise mechanisms that drive these chemical changes must be known.

The goal of this thesis was to examine in detail some of the different

photochemical processes that can occur for environmental pollutants upon exposure to

sunlight under aqueous conditions. Three different mechanistic systems were studied,

consisting of both direct and indirect photochemical reactions. In the first and third

sections, our goal was to take a closer look at the source of pollutants to see if

photochemical transformations could help explain the presence of specific compounds

in the aquatic environment. In the second section, we performed a more fundamental

study to understand indirect photooxidation reactions with the goal of aiding in

transformation rate predictions for some pollutant classes.

The first part of this thesis examines the phototransformation of hydroxylated

polybrominated diphenyl ethers (OH-PBDEs), which are known to be both natural

Page 7: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 2  

products as well as transformation products of flame retardant chemicals. The

reactivity of some of these compounds via direct photolysis was investigated, and a

number of transformation products were identified and quantified. The most

concerning transformation products of OH-PBDEs were polybrominated dibenzo-p-

dioxins (PBDDs), which were formed by a photochemical ring closure reaction that

all of the studied compounds were found to undergo with differing efficiencies. The

OH-PBDE congener, 6-OH-PBDE 99, which is known to be a natural product,

generated the most PBDD with a yield of 7%. Another congener, 6’-OH-PBDE 118,

which is thought to only be a transformation product of other anthropogenic

compounds was capable of generating 2,3,7,8-PBDD, one of the most toxic PBDD

congeners, but with yields of only 0.5%. The results of this study further support the

findings of others that most of the PBDDs found in marine environments, more

specifically the Baltic Sea, are derived from natural sources through a variety of

production mechanisms.

The second part of this thesis was devoted to studying the indirect

photochemical oxidation of two sets of model pollutants. The reaction mechanism for

the oxidation of anilines and phenols by a model triplet state oxidant, Methylene Blue

(MB) was investigated in aqueous solution. Using transient absorption spectroscopy,

the reaction rate constants were determined for the oxidative reaction between MB

and series of substituted anilines and phenols. The goal of this project was to

determine an empirical relationship between the reaction free energy and reaction rate

constant for each set of model pollutants to aid in the understanding of the mechanism

as well as fate prediction of these compounds in the environment. The anilines, which

reacted with MB by a one-electron transfer mechanism, were observed to react with

rate constants ranging from 1.06x107 M-1 s-1 to 4.85x109 M-1 s-1. These reaction rate

constants were found to correlate well to the reaction free energy when fit using a

Page 8: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  3  

Sandros-Boltzmann distribution model. The reaction between MB and phenols was

found to occur by a proton concerted electron transfer mechanism. Evidence for this

reaction mechanism over a simple electron transfer came from spectroscopic data,

which suggested the immediate formation of a protonated MB species following

reaction. Additional kinetic isotope effect experiments further demonstrated that the

OH bond strength of the phenols directly affected the observed reaction rate

constants. The reaction rate constants for substituted phenols could be predicted with

high accuracy by a linear free energy relationship that depended on the bond

disassociation free energy and the pKa.

In the third section of this thesis, the mechanism of the photochemical

formation of 1-chloropyrene from pyrene and chloride in aqueous solution was

investigated. The motivation for this work came from the observation that the

concentration of 1-chloropyrene found in marine sediments correlated with the water

salinity, thus suggesting that the formation from pyrene was occurring in situ. A

combination of steady state photolysis experiments and transient absorption

measurements were carried out to study this reaction mechanism. Our results suggest

1-chloropyrene is formed by a nucleophilic attack of chloride on the pyrene radical

cation, which is generated by the oxidation of photoexcited pyrene by O2. The

nucleophilic behavior of chloride was demonstrated by performing photolysis

experiments in the presence of chloride and other nucleophiles. When pyrene was

irradiated in the presence of equal concentrations of bromide and chloride, 1-

bromopyrene was observed to form at approximately ten times the rate of 1-

chloropyrene. Based on the relative nucleophilicities of bromide and chloride these

results supports a nucleophilic mechanism. The addition of electron donors, which

were observed to quench the pyrene radical cation signal in transient absorption

Page 9: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 4  

experiments, also eliminated the production of 1-chloropyrene during steady-state

photolysis.

Overall, this thesis added to the understanding of three different mechanistic

systems relevant to the fate of organic pollutants in aquatic environments. The results

of this work demonstrate that laboratory-scale experiments are important in

explaining the presence of specific environmental pollutants based on transformations

occurring within the environment. Additionally, our investigation of the mechanisms

of indirect photooxidation and modeling of the observed reaction rate constants may

be crucial in future attempts to create pollutant fate prediction tools.

Page 10: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  5  

Zusammenfassung

Die Freisetzung von organischen Schadstoffen in der Umwelt ist ein globales

Problem und kann negative Auswirkungen auf die Umwelt und auf den Menschen

haben. Um die Risiken für Mensch und Umwelt ausgehend von der Verwendung und

der Beseitigung von Pharmazeutika, Pestiziden und anderen industriell produzierten

Chemikalien genau abschätzen zu können, sind gute Kenntnisse über deren Verbleib

in der Umwelt von Nöten. Um das von einer chemischen Verbindung ausgehende

Risiko zu erfassen, müssen wir zuerst die wichtigen Abbauprozesse ermitteln und

abklären ob umweltschädliche Transformationsprodukte gebildet werden.. Für viele

Umweltschadstoffe stellen die Aussetzung zum Sonnenlicht und Interaktionen mit

reaktiven photochemisch produzierten Zwischenprodukten die wichtigsten

Umwandlungsprozesse dar. Viele mögliche Photoreaktionen können den Abbau von

organische Substanzen beeinflussen. Um das Verhalten eines Schadstoffes in der

Umwelt vorhersagen zu können, ist ein detailliertes Verständnis der chemischen

Mechanismen dieser Photoreaktionen von Nöten.

Diese Doktorarbeit verfolgte das Ziel einige oxidative Prozesse von

Umweltschadstoffen zu untersuchen, die im Wasser durch Sonneneinstrahlung

erfolgen. Die drei untersuchten mechanistischen Systeme schlossen sowohl direkte als

auch indirekte Photoreaktionen ein. Im ersten sowie im dritten Teil dieser Studie

haben wir das Ziel verfolgt, die Herkunft von Schadstoffen genauer zu untersuchen

um zu sehen ob Photoreaktionen das Vorkommen dieser Schadstoffe erklären können.

Im dritten Fall wurde eine eher fundamentale Studie durchgeführt um indirekte

photooxidative Reaktionen genauer zu verstehen und damit Transformationsraten für

verschiedene Substanzklassen vorhersagen zu können.

Page 11: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 6  

Im ersten Teil dieser Doktorarbeit geht es um die Phototransformation von

hydroxylierten polybromierten Diphenylethern (OH-PBDEs), welche als natürliche

Produkte aber auch als Transformationsprodukte von Flammschutzmitteln in der

Umwelt vorkommen. Für einige dieser Verbindungen wurde die Reaktivität durch

direkte Lichtabsorption untersucht und Transformationsprodukte wurden identifiziert

und quantifiziert. Die bedenklichsten Transformationsprodukte der OH-PBDEs waren

die polybromierten dibenzo-p-Dioxine (PBDDs), welche durch eine photochemische

Ringschliessung gebildet wurden. Alle untersuchten Verbindungen gingen diese

Ringschliessung ein, mit unterschiedlicher Effizienz. Das OH-PBDE Kongener, 6-

OH-PBDE 99, ein Naturprodukt, bildete mit einer Ausbeute von 7% am meisten

PBDD. Ein anderes Kongener, 6’-OH-PBDE 118, welches vermutlich ein

Transformationsprodukt anthropogener Verbindungen ist, bildete 2,3,7,8-PBDD,

eines der giftigsten PBDD Kongeneren, allerdings mit einer Ausbeute von nur 0.5%.

Die Resultate dieser Studie bestätigen andere Forschungsergebnisse, welche zeigten

dass die meisten PBDDs in der marinen Umwelt, konkret im baltischen Meer, aus

natürlichen Quellen stammen und durch verschiedene Mechanismen gebildet werden

können.

Im zweiten Teil dieser Doktorarbeit wurde die indirekte photochemische

Oxidation von zwei Modellschadstoffklassen untersucht. Die Reaktionsmechanismen

für die Oxidation von Anilinen und Phenolen wurde anhand eines Modell-Triplet-

Oxidationsmittel, Methylenblau (MB) untersucht. Die Geschwindigkeitskonstanten

für die oxidative Reaktion zwischen MB und einer Reihe von substituierten Anilinen

und Phenolen wurden mittels transienter Absorptionsspektroskopie ermittelt. Das Ziel

dieses Projekts war es, eine empirische Beziehung zwischen den freien

Bindungsenthalpien und den Geschwindigkeitskonstanten für jede der zwei

Modellschadstoffklassem zu ermitteln und damit zum Verständnis des

Page 12: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  7  

Reaktionsmechanismus und dem Verbleib dieser Verbindungsklassen in der Umwelt

beizutragen. Für die Aniline, welche mit MB mit einem Ein-

Elektronentransfermechanismus reagierten, wurden Reaktionskonstanten zwischen

1.06x107 M-1 s-1 und 4.85x109 M-1 s-1 beobachtet. Diese Geschwindigkeitskonstanten

korrelierten gut mit der freien Bindungsenergie wenn die Daten an ein Sandros-

Boltzmann Verteilungsmodell angepasst wurden. Die Reaktion zwischen MB und den

Phenolen folgte einem Protonen-gekoppeltem Elektronentransfermechanismus. Ein

einfacher Elektrontransfermechanismus konnte aufgrund spektroskopischer Daten

ausgeschlossen werden, da diese auf eine sofortige Bildung von protoniertem MB

nach der Reaktion hinweisen. Weitere Experimente zum kinetischen Isotopeneffekt

zeigten, dass die OH-Bindungsstärke der Phenole die beobachteten

Geschwindigkeitskonstanten direkt beeinflussten. Die Geschwindigkeitskonstanten

für die substituierten Phenole konnten mit hoher Genauigkeit durch eine lineare freie

Energie-Beziehung, abhängig von der freien Bindungsenergie und dem pKa,

vorhergesagt werden.

Im dritten Teil dieser Doktorarbeit wurde der Mechanismus der

photochemischen Bildung von 1-Chloropyren aus Pyren und Chlorid in wässriger

Lösung untersucht. Diese Forschungsarbeit wurde motiviert durch die Beobachtung,

dass ein Zusammenhang zwischen der Konzentration von 1-Chloropyren in marinen

Sedimenten und der Wassersalinität besteht, welcher auf in-situ Bildung aus Pyren

hindeutete. Der Mechanismus wurde mit einer Kombination aus stationärer

Bestrahlung und transienten Absorptionsexperimenten untersucht. Unsere Resultate

zeigen, dass 1-Chloropyren durch den nukleophilen Angriff von Chlorid auf das

Pyrenradikalkation gebildet wird, welches wiederum durch die Oxidation von

lichtangeregtem Pyren durch O2 entstand. Dass sich Chlorid als Nukleophil verhält

wurde in Photolyseexperimenten mit Chlorid und anderen Nukleophilen demonstriert.

Page 13: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 8  

Wenn Pyren in der Gegenwart von gleichen Konzentrationen an Chlorid- und

Bromidionen photolysiert wird, entsteht ungefähr zehn mal mehr 1-Bromopyren als 1-

Chloropyren. Die relativen Nukleophilitäten von Chlorid und Bromid lassen auf einen

nukleophilen Mechanismus schliessen. Ausserdem wurde gezeigt, dass die Zugabe

von Elektronendonoren, welche in den transienten Absorptionsexperimenten das

Pyrenradikalsignal abgeschwächt haben, auch die Produktion von 1-Chloropyren

während der stationären Bestrahlung verhinderten.

Zusammenfassend hat diese Doktorarbeit zum Verständnis von drei

verschiedenen mechanistischen Systemen beigetragen, welche alle für das

Verständnis des Verbleibs von organischen Schadstoffen in Gewässern relevant sind.

Die Resultate dieser Arbeit zeigen, dass Experimente im Labor wichtig sind um das

Vorkommen von gewissen Umweltschadstoffen zu erklären, welche durch

Transformationen in der Umwelt entstanden sind. Zusätzlich kann unsere

Untersuchung von indirekten photooxidativen Mechanismen und die Modellierung

von beobachteten Geschwindigkeitskonstanten entscheidend sein für zukünftige

Tools, die den Verbleib von Schadstoffen in der Umwelt versuchen vorherzusagen.

Page 14: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  9  

Chapter 1

Introduction

Page 15: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 10  

Introduction

For nearly two centuries, our ability to produce and use synthetic organic

chemicals has played a vital role in improving the quality of life for billions of people

worldwide. Thanks largely to the discovery of pharmaceuticals, fertilizers, and

countless other chemical technologies, it is possible to feed and maintain the health of

a growing world population. Unfortunately, many of the same chemicals we have

used to improve our lives have had unintended and in some cases disastrous

consequences on human and environmental health. A prominent and early example

was the use of 1,1,1-trichloro-2,2-bis(4-chlorophenyl)ethane, or DDT. Most people

today only know DDT as a dangerous environmental pollutant, but in the 1940’s its

powerful abilities as an insecticide are credited, in part, for eliminating the

transmission of malaria by mosquitoes in the United States during the 1950s1. DDT

was once so highly regarded, the Swiss chemist Paul Müller received the 1948 Nobel

prize in Physiology or Medicine for the discovery of DDTs insecticidal properties.

This achievement however became greatly overshadowed by the detrimental effect

that DDT use had on wildlife and human health, which has led to its eventual ban in

most developed countries since the 1970s1. The problems resulting from DDT and

many other pollutants that end up in the environment make it necessary for us to

apply the same science of chemistry to understand the effects, persistence, and fate of

these chemicals in the natural world.

Pollutants enter the environment in a wide range of ways, and their release can

be both intentional, such as the spraying of pesticides and herbicides for agricultural

purposes, or unintentional, as is the case for the thousands of micropollutants that are

released through wastewater treatment and modern industrial practices. Once in the

Page 16: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  11  

environment, these compounds partition into various compartments depending on

their chemical properties, including the air, water, soil, and biota. Pollutants that end

up in aquatic environments in either a free or particle-bound state are subjected to a

unique set of conditions that determines how and by what mechanisms they may

transform. In surface waters, an important source of transformation for many

pollutants is photolysis via exposure to sunlight. These light-initiated reactions are

typically divided into two categories, direct and indirect photolysis.

Fundamentals of light

Before discussing the basics of photochemistry, it is useful to first introduce

some fundamentals about the unifying and most important component of these

reactions, light. At the most basic level, light, or electromagnetic radiation, is the

carrier of energy between the electrons found in atoms and molecules. This transfer of

energy is carried out by massless fundamental particles known as photons, which

carry discrete packets, or quanta of energy. Because of the dual particle-wave nature

of photons2, it is convenient to describe their energy according to the frequency with

which their electromagnetic field oscillates through space:

(1)

where E is the energy carried by the photon in J, h is Plank’s constant, 6.626x1023 J

s), and n is the frequency of light in units of s-1 . While the frequency of light

determines its energy, it is often more common in photochemistry to discuss the

energy of light in terms of its wavelength: l:

(2)

where c is the speed of light in a vacuum. When we speak about the relative energies

of light in terms of wavelength, higher energy photons have a shorter wavelength (e.g.

ultraviolet light) than lower energy photons (infrared). For reference, the human eye is

E = hν

λ =cν

Page 17: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 12  

sensitive to photons with wavelengths between 400 nm and 700 nm, which is known

as the visible spectrum. In environmental photochemistry, the most important light

source is the sun. At the surface of the earth, the solar spectrum contains wavelengths

of light ranging from approximately 290 to 2500 nm, with the daytime peak emission

wavelength being around 500 nm.

Direct photolysis

The first and most simple requirement in a direct photochemical reaction is

that the compound of interest must absorb a photon. The Grotthus-Draper law, or the

first law of photochemistry, states that only light that is absorbed by a system is

capable of producing chemical changes. The Stark-Einstein law, or second law of

photochemistry, states that every molecule participating in a photochemical reaction

must absorb one quantum of light. This does not mean that for every photon absorbed

only one reaction may take place, however. In come cases, the chemical change

brought on by the absorption of one photon can lead to a chain of secondary non-

photochemical reactions, as is the case for some radical processes. The important

distinction is that only one photochemical reaction takes place, and that the

subsequent reactions are thermal. We can describe a simple, dilute solution where

molecule R undergoes a direct photochemical reaction to form P using equation 3:

(3)

The first-order reaction rate constant kdeg for the degradation of P determines how

quickly R will be transformed, therefore it is desirable to determine this rate constant

for the purposes of predicting pollutant transformation rates in the environment. kdeg

can be expressed as:

R Ph!+kdeg

Page 18: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  13  

(4)

where e is the molar extinction coefficient in units M-1cm-1, which describes the

efficiency with which a compound absorbs light at each wavelength, and I is the

irradiance of the light source in units millieinsteins cm-2 s-1. These two terms summed

up over all the relevant wavelengths determine the rate that a compound absorbs light,

while the remaining term, , or the degradation quantum yield, informs us about

how efficient the absorbed photons are at inducing chemical change in the molecule.

is simply the ratio of the number of chemical reactions induced divided by the

number of photons absorbed. The value of can range from nearly 0 for

molecules that are highly efficient at releasing their energy through physical

processes, such as re-emission of a photon in fluorescence, up to values greater than

1000 for cases mentioned earlier where a single photochemical reaction leads to

further thermal reactions. In dilute aqueous solutions, however, the value of

typically ranges between 0 and 1 since the likelihood of chain reactions is rather low.

Direct photolysis has been known to play a role in the removal and

transformation of a wide range of organic pollutants capable of absorbing sunlight.

After light absorption, these photoexcited molecules may undergo a wide variety of

reactions yielding new products. Scheme 2 highlights a few examples of direct

photochemical transformations seen in aqueous environments. One of the simplest

phototransformation pathways is photoionization. In a photoionization reaction, an

electron in a molecule acquires enough energy by absorbing a photon that it can

simply be ejected out of its orbital and taken up by the surrounding solvent. This

reaction has been shown to play a role in the degradation of aniline and its

derivatives3,4, indole-containing molecules5,6, and other environmental pollutants.

kdeg = Φdeg ελIλλ∑

Φdeg

Φdeg

Φdeg

Φdeg

Page 19: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 14  

Once ionized, the resulting radical cations of these molecules undergo further

transformation if they do not encounter a suitable electron donor in solution.

Scheme 1- Selected examples of direct photochemical reactions

With the exception of fluorine, carbon-halogen bonds are relatively weak

compared in comparison to C-H bonds in aromatic compounds. As an initial step in

transformation, halogenated aromatic compounds often undergo C-X bond cleavage,

which can be followed by hydroxylation7,8, hydrogenation9, and various

rearrangements10-12.

While direct photochemical reactions often lead to the removal of pollutants,

there exist a number of examples where direct photolysis leads to the production of

molecules with more troublesome properties than the parent compound. A good

example of this is the photochemical formation of polychlorinated p-dibenzodioxins,

which are known toxins13-16 and heavily regulated as environmental pollutants,

NH2

h!NH2

+ eaq-

OCl

Cl Cl

OHO

Cl ClO

h! + HCl

OHCl h!

OH

OH2O + + Cl-

Cl

H2O +h!

OH

+ HCl

Page 20: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  15  

following a photochemical ring closure reaction by triclosan17-21 and other

halogenated diphenyl ether derivatives22. Another interesting observation is that some

molecules that undergo direct photolysis in water may revert back to their parent

compounds in the dark. The steroid trenbolone acetate has been shown23,24 to undergo

photochemical hydration reaction followed by a thermal dehydration in aqueous

solution. This can lead to a diurnal variation in the observed concentration, and not

accounting for this regeneration results in an underestimation of the lifetime of this

compound in aquatic environments. In these cases, simply knowing the rate of direct

photolysis for a pollutant fails to accurately capture its true environmental impact, and

emphasizes that we must also study and account for the transformation products of

environmental contaminants.

Indirect photolysis

Given the limited range of UV wavelengths in the solar spectrum at earth’s

surface, it comes as no surprise that many organic pollutants cannot absorb sunlight,

and thus are not transformed by direct photolysis. Oftentimes these compounds may

still be transformed by sunlight in an alternative way known as indirect photolysis. In

an indirect photoreaction, another molecule, called a sensitizer, first absorbs light and

either transfers energy to or reacts with the target molecule. A general expression for

this type reaction is shown in equation 5:

(5)

Additionally, sensitizers may also transfer their energy to or generate an intermediate

molecule, which can go on to react further with pollutants. In aquatic environments

one of the dominant light absorbers and the most common sensitizer is chromophoric

Sh!

S* + Rksens S + R* or P

Page 21: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 16  

dissolved organic matter25,26 (CDOM). CDOM can sensitize the formation of a

number of reactive, oxygen based intermediates collectively known as reactive

oxygen species (ROS).

Scheme 2- Examples of indirect photochemical oxidations

Some of the more important ROS species generated by the photolysis of CDOM

include singlet oxygen (1O2), hydrogen peroxide (H2O2), and hydroxyl radical (HO!).

1O2 is simply an excited state of O2, differing from its ground state only by the spin

direction of one electron, and is a selective but powerful oxidant for some classes of

molecules. In the aqueous phase, 1O2 is short-lived and thus present in sunlit aquatic

environments in the pM concentration range27. Because of the unique electronic

structure of 1O2, it is particularly well suited to participating in ring-forming addition

reactions, and has been shown to play an important role in the aqueous degradation of

anthracene28 via a formal 2+4 cycloaddition29 and the amino acid tryptophan30 via a

formal 2+2 cycloaddition. H2O2 is a less reactive, but much longer-lived ROS, and

can be found at significantly higher concentrations than 1O2 in aquatic environments.

HO! is the most reactive of all the ROS known to be generated from CDOM, reacting

unselectively at diffusion-controlled rates with most organic compounds. Because of

+ 1O2OO

+ HO + HOH

ClHN

O

N+ 3CDOM*

ClHN

O

N+ CDOM

Page 22: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  17  

this high reactivity, HO! plays an important role in the degradation of some of the

most recalcitrant environmental pollutants31,32.

Much of the indirect photochemical behavior seen by CDOM can be explained

by the presence of excited triplet state molecules33-36 (3CDOM*) that are generated

following irradiation. In addition to being a source of ROS, 3CDOM* can also act as

an oxidant for environmental pollutants. Following excitation, the molecules that

make up 3CDOM* become stronger oxidants, and thus may act as one-electron

acceptors for a wide range of pollutants. These types of reactions are known as

photoinduced electron transfers. Reactions of this type by 3CDOM* have been shown

to be important in degrading pharmacuticals37, phenols36,38, phenylurea herbicides37,

as well as other molecules35 naturally present in the environment. The main

components of CDOM thought to be responsible for the reactivity of 3CDOM* are

aromatic ketones39. Several studies have shown that aromatic ketones in their first

excited triplet state show high reactivity for one-electron oxidation reactions for

compounds such as phenols34. Depending on the exact structure and reduction

potential of these ketones, they may also react by acting as hydrogen atom acceptors

for phenols and other labile molecules36,40,41.

Goals and thesis overview

The goal of this thesis was to investigate the role of various direct and indirect

photochemical processes for different organic micropollutants in the aqueous phase.

The scope of each chapter is different, but each involves the detailed study of a photo-

oxidative mechanism related to known environmental pollutants. In Chapter 2, the

direct photolysis of hydroxylated polybrominated diphenyl ethers (OH-PBDEs) was

studied. OH-PBDEs are known transformation products42-46 of a class of flame

retardants known as polybrominated diphenyl ethers (PBDEs) which have now been

Page 23: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 18  

banned due to their widespread environmental presence and persistence47,48. In

addition to being transformation products of PBDEs, OH-PBDEs are also known to

be natural products49-51. Our main objective was to study the photochemical ring

closure of OH-PBDEs, which leads to the formation of polybrominated p-

dibenzodioxins (PBDDs), a pollutant of even greater concern. We quantified the

production of PBDDs and other phototransformation products of selected OH-PBDEs

that were known to be transformation products of PBDEs. The goal of this work was

to aid in determining the anthropogenic contribution to the observed concentration

levels of PBDDs in marine environments.

In Chapter 3, the indirect photochemical oxidation mechanisms for two sets of

model pollutants was investigated with a triplet excited state oxidant, methylene blue

(MB). Motivated by the observation that many environmental micropollutants seem to

be oxidized by one-electron transfer reactions with 3CDOM34-36,38,52, we attempted to

determine models capable of predicting bimolecular oxidation rate constants using

both substituted anilines and phenols. Using transient absorption spectroscopy, the

quenching of triplet MB by the model compounds was directly observed in order to

determine reaction rate constants. The mechanism of photooxidation for anilines was

found to be a one electron transfer from the aniline to MB, while phenols were

observed to react by proton concerted electron transfer (PCET). In both cases, an

adequate correlation between the oxidation reaction free energy and the bimolecular

rate constant were found. For the electron transfer driven oxidation of anilines,

bimolecular reaction rate constants were found to fit well to a Sandros-Boltzmann

distribution, which has only recently been employed as an empirical model37,53 to

describe photoinduced electron transfer reactions. For phenols, a good correlation

between the phenolic OH bond dissociation free energy (BDFE) and the reaction rates

was found. In plots of the reaction rate vs. BDFE however, a systematic deviation

Page 24: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  19  

proportional to the phenolic pKa was observed. Once factored into our model, the

observed PCET reaction rates constants were in good agreement with predicted rates.

To supplement the lack of reliable experimental aqueous oxidation potentials for

anilines, which are needed to calculate electron transfer free energies, and to aid in

future predictive models, the oxidation potentials for the substituted anilines were

calculated in silico. In addition, the BDFEs of phenols were also computed.

In chapter 4, the focus was again on another direct photochemical oxidation

process, the mechanism of the photochemical chlorination of pyrene (PYR). It has

been shown recently54,55 that the concentration of 1-chloropyrene (Cl-PYR) and other

chlorinated polycyclic aromatic hydrocarbons in marine sediments correlates strongly

with water salinity, furthermore Cl-PYR was shown to be a transformation product of

the photolysis of PYR in artificial seawater. Our goal was to test the hypothesis that

Cl-PYR is formed in marine environments by chloride acting as a nucleophile and

attacking the pyrene radical cation of (PYR!+) that is formed by the oxidation of

photoexcited PYR. Our experiments showed that the formation of Cl-PYR depends on

the presence of O2, which is responsible for oxidizing the excited PYR to form the

intermediate radical cation, and that Cl-PYR formation can also be suppressed by the

addition of an electron donor (N3-). The role of PYR!+ is further supported by transient

absorption observations that indicate the suppression of PYR!+ in oxygen-free

conditions and rapid quenching by N3-. These and other experiments described in

chapter 4 support our hypothesis of Cl-PYR formation, and provide further evidence

of the in situ formation of chlorinated PAHs in some marine environments.

Page 25: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 20  

References

(1) DDT and Its Derivatives (Environmental Health Criteria), World Health Organization, 1979. (2) Turro, N. J.; Ramamurthy, V.; Scaiano, J. C. Principles of Molecular Photochemistry: An Introduction; University Science Books, 2009. (3) Saito, F.; Tobita, S.; Shizuka, H. Photoionization of aniline in aqueous solution and its photolysis in cyclohexane. Journal of the Chemical Society-Faraday Transactions 1996, 92, 4177-4185. (4) Saito, F.; Tobita, S.; Shizuka, H. Photoionization mechanism of aniline derivatives in aqueous solution studied by laser flash photolysis. Journal of Photochemistry and Photobiology a-Chemistry 1997, 106, 119-126. (5) Katoh, R. Dependence of photoionization quantum yield of indole and tryptophan in water on excitation wavelength. Journal of Photochemistry and Photobiology a-Chemistry 2007, 189, 211-217. (6) Tsentalovich, Y. P.; Snytnikova, O. A.; Sagdeev, R. Z. Properties of excited states of aqueous tryptophan. Journal of Photochemistry and Photobiology a-Chemistry 2004, 162, 371-379. (7) Choudhry, G. G.; Vanderwielen, F. W. M.; Webster, G. R. B.; Hutzinger, O. PHOTOCHEMISTRY OF HALOGENATED BENZENE-DERIVATIVES .6. PHOTOREACTIONS OF TETRACHLOROPHENOLS AND PENTACHLOROPHENOLS IN WATER ACETONITRILE MIXTURES. Canadian Journal of Chemistry-Revue Canadienne De Chimie 1985, 63, 469-475. (8) Rayne, S.; Wan, P.; Ikonomou, M. Photochemistry of a major commercial polybrominated diphenyl ether flame retardant congener: 2,2′,4,4′,5,5′-Hexabromodiphenyl ether (BDE153). Environment International 2006, 32, 575-585. (9) Kliegman, S.; Eustis, S. N.; Arnold, W. A.; McNeill, K. Experimental and Theoretical Insights into the Involvement of Radicals in Triclosan Phototransformation. Environmental Science & Technology, 47, 6756-6763. (10) . (11) Boule, P.; Guyon, C.; Lemaire, J. PHOTOCHEMISTRY AND ENVIRONMENT .4. PHOTOCHEMICAL BEHAVIOR OF MONOCHLOROPHENOLS IN DILUTE AQUEOUS-SOLUTION. Chemosphere 1982, 11, 1179-1188. (12) Rayne, S.; Forest, K.; Friesen, K. J. Mechanistic aspects regarding the direct aqueous environmental photochemistry of phenol and its simple halogenated derivatives. A review. Environment International 2009, 35, 425-437. (13) Truce, W. E.; Kreider, E. M.; Brand, W. W. In Organic Reactions; John Wiley & Sons, Inc.: 2011. (14) Ao, K.; Suzuki, T.; Murai, H.; Matsumoto, M.; Nagai, H.; Miyamoto, Y.; Tohyama, C.; Nohara, K. Comparison of immunotoxicity among tetrachloro-, pentachloro-, tetrabromo- and pentabromo-dibenzo-p-dioxins in mice. Toxicology 2009, 256, 25-31. (15) Birnbaum, L. S.; Staskal, D. F.; Diliberto, J. J. Health effects of polybrominated dibenzo-p-dioxins (PBDDs) and dibenzofurans (PBDFs). Environment International 2003, 29, 855-860. (16) Van den Berg, M.; Birnbaum, L. S.; Denison, M.; De Vito, M.; Farland, W.; Feeley, M.; Fiedler, H.; Hakansson, H.; Hanberg, A.; Haws, L.; Rose, M.; Safe, S.; Schrenk, D.; Tohyama, C.; Tritscher, A.; Tuomisto, J.; Tysklind, M.;

Page 26: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  21  

Walker, N.; Peterson, R. E. The 2005 World Health Organization Reevaluation of Human and Mammalian Toxic Equivalency Factors for Dioxins and Dioxin-Like Compounds. Toxicological Sciences 2006, 93, 223-241. (17) Weber, R.; Tysklind, M.; Gaus, C. Dioxin - contemporary and future challenges of historical legacies. Environmental Science and Pollution Research 2008, 15, 96-100. (18) Kanetoshi, A.; Ogawa, H.; Katsura, E.; Kaneshima, H. Chlorination of irgasan DP300 and formation of dioxins from its chlorinated derivatives. Journal of Chromatography A 1987, 389, 139-153. (19) Buth, J. M.; Grandbois, M.; Vikesland, P. J.; McNeill, K.; Arnold, W. A. Aquatic photochemistry of chlorinated triclosan derivatives: Potential source of polychlorodibenzo-P-dioxins. Environmental Toxicology and Chemistry 2009, 28, 2555-2563. (20) Latch, D. E.; Packer, J. L.; Arnold, W. A.; McNeill, K. Photochemical conversion of triclosan to 2,8-dichlorodibenzo-p-dioxin in aqueous solution. Journal of Photochemistry and Photobiology A: Chemistry 2003, 158, 63-66. (21) Latch, D. E.; Packer, J. L.; Stender, B. L.; VanOverbeke, J.; Arnold, W. A.; McNeill, K. Aqueous photochemistry of triclosan: Formation of 2,4-dichlorophenol, 2,8-dichlorodibenzo-p-dioxin, and oligomerization products. Environmental Toxicology and Chemistry 2005, 24, 517-525. (22) Sanchez-Prado, L.; Llompart, M.; Lores, M.; García-Jares, C.; Bayona, J. M.; Cela, R. Monitoring the photochemical degradation of triclosan in wastewater by UV light and sunlight using solid-phase microextraction. Chemosphere 2006, 65, 1338-1347. (23) Bastos, P. M.; Eriksson, J.; Bergman, Å. Photochemical decomposition of dissolved hydroxylated polybrominated diphenyl ethers under various aqueous conditions. Chemosphere 2009, 77, 791-797. (24) 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, M. D.; Cwiertny, D. M. Product-to-Parent Reversion of Trenbolone: Unrecognized Risks for Endocrine Disruption. Science 2013, 342, 347-351. (25) 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 Photo-Transformation Products of Trenbolone Acetate Metabolites. Environmental Science & Technology, 47, 5031-5041. (26) Blough, N. V.; Zafiriou, O. C.; Bonilla, J. OPTICAL-ABSORPTION SPECTRA OF WATERS FROM THE ORINOCO RIVER OUTFLOW - TERRESTRIAL INPUT OF COLORED ORGANIC-MATTER TO THE CARIBBEAN. Journal of Geophysical Research-Oceans 1993, 98, 2271-2278. (27) Green, S. A.; Blough, N. V. OPTICAL-ABSORPTION AND FLUORESCENCE PROPERTIES OF CHROMOPHORIC DISSOLVED ORGANIC-MATTER IN NATURAL-WATERS. Limnol. Oceanogr. 1994, 39, 1903-1916. (28) Haag, W. R.; Hoigne, J. Singlet oxygen in surface waters. 3. Photochemical formation and steady-state concentrations in various types of waters. Environmental Science & Technology 1986, 20, 341-348. (29) Zafiriou, O. C. MARINE ORGANIC-PHOTOCHEMISTRY PREVIEWED. Marine Chemistry 1977, 5, 497-522. (30) Corey, E. J.; Taylor, W. C. A Study of the Peroxidation of Organic Compounds by Externally Generated Singlet Oxygen Molecules. Journal of the American Chemical Society 1964, 86, 3881-3882.

Page 27: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 22  

(31) Boreen, A. L.; Edhlund, B. L.; Cotner, J. B.; McNeill, K. Indirect Photodegradation of Dissolved Free Amino Acids: The Contribution of Singlet Oxygen and the Differential Reactivity of DOM from Various Sources. Environmental Science & Technology 2008, 42, 5492-5498. (32) Lam, M. W.; Tantuco, K.; Mabury, S. A. PhotoFate:   A New Approach in Accounting for the Contribution of Indirect Photolysis of Pesticides and Pharmaceuticals in Surface Waters. Environmental Science & Technology 2003, 37, 899-907. (33) Stangroom, S. J.; MacLeod, C. L.; Lester, J. N. Photosensitized transformation of the herbicide 4-chloro-2-methylphenoxy acetic acid (MCPA) in water. Water Research 1998, 32, 623-632. (34) Canonica, S. Oxidation of Aquatic Organic Contaminants Induced by Excited Triplet States. CHIMIA International Journal for Chemistry 2007, 61, 641-644. (35) Canonica, S.; Hellrung, B.; Müller, P.; Wirz, J. Aqueous Oxidation of Phenylurea Herbicides by Triplet Aromatic Ketones. Environmental Science & Technology 2006, 40, 6636-6641. (36) Canonica, S.; Hellrung, B.; Wirz, J. Oxidation of Phenols by Triplet Aromatic Ketones in Aqueous Solution. The Journal of Physical Chemistry A 2000, 104, 1226-1232. (37) Canonica, S.; Jans, U.; Stemmler, K.; Hoigne, J. Transformation Kinetics of Phenols in Water: Photosensitization by Dissolved Natural Organic Material and Aromatic Ketones. Environmental Science & Technology 1995, 29, 1822-1831. (38) Boreen, A. L.; Arnold, W. A.; McNeill, K. Triplet-Sensitized Photodegradation of Sulfa Drugs Containing Six-Membered Heterocyclic Groups:   Identification of an SO2 Extrusion Photoproduct. Environmental Science & Technology 2005, 39, 3630-3638. (39) McNally, A. M.; Moody, E. C.; McNeill, K. Kinetics and mechanism of the sensitized photodegradation of lignin model compounds. Photochemical & Photobiological Sciences 2005, 4, 268-274. (40) Das, P. K.; Encinas, M. V.; Scaiano, J. C. Laser flash photolysis study of the reactions of carbonyl triplets with phenols and photochemistry of p-hydroxypropiophenone. Journal of the American Chemical Society 1981, 103, 4154-4162. (41) Lathioor, E. C.; Leigh, W. J. Bimolecular Hydrogen Abstraction from Phenols by Aromatic Ketone Triplets†. Photochemistry and Photobiology 2006, 82, 291-300. (42) Leigh, W. J.; Lathioor, E. C.; St. Pierre, M. J. Photoinduced Hydrogen Abstraction from Phenols by Aromatic Ketones. A New Mechanism for Hydrogen Abstraction by Carbonyl n,π* and π,π* Triplets. Journal of the American Chemical Society 1996, 118, 12339-12348. (43) Athanasiadou, M.; Cuadra, S. N.; Marsh, G.; Bergman, Å.; Jakobsson, K. Polybrominated Diphenyl Ethers (PBDEs) and Bioaccumulative Hydroxylated PBDE Metabolites in Young Humans from Managua, Nicaragua. Environ Health Perspect 2007, 116. (44) Bergman, A.; Athanasiadou, M.; Faeldt, E.; Jakobsson, K. Hydroxylated PBDE metabolites in human blood. Organohalogen Compd. 2006, 68, 635-638.

Page 28: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  23  

(45) Raff, J. D.; Hites, R. A. Gas-Phase Reactions of Brominated Diphenyl Ethers with OH Radicals. The Journal of Physical Chemistry A 2006, 110, 10783-10792. (46) Ueno, D.; Darling, C.; Alaee, M.; Pacepavicius, G.; Teixeira, C.; Campbell, L.; Letcher, R. J.; Bergman, Å.; Marsh, G.; Muir, D. Hydroxylated Polybrominated Diphenyl Ethers (OH-PBDEs) in the Abiotic Environment: Surface Water and Precipitation from Ontario, Canada. Environmental Science & Technology 2008, 42, 1657-1664. (47) Wan, Y.; Choi, K.; Kim, S.; Ji, K.; Chang, H.; Wiseman, S.; Jones, P. D.; Khim, J. S.; Park, S.; Park, J.; Lam, M. H. W.; Giesy, J. P. Hydroxylated Polybrominated Diphenyl Ethers and Bisphenol A in Pregnant Women and Their Matching Fetuses: Placental Transfer and Potential Risks. Environ. Sci. Technol., 44, 5233-5239. (48) D'Silva, K.; Fernandes, A.; Rose, M. Brominated Organic Micropollutants—Igniting the Flame Retardant Issue. Critical Reviews in Environmental Science and Technology 2004, 34, 141-207. (49) de Wit, C. A. An overview of brominated flame retardants in the environment. Chemosphere 2002, 46, 583-624. (50) Malmvärn, A.; Marsh, G.; Kautsky, L.; Athanasiadou, M.; Bergman, Å.; Asplund, L. Hydroxylated and Methoxylated Brominated Diphenyl Ethers in the Red Algae Ceramium tenuicorne and Blue Mussels from the Baltic Sea. Environmental Science & Technology 2005, 39, 2990-2997. (51) Malmvärn, A.; Zebühr, Y.; Kautsky, L.; Bergman, Å.; Asplund, L. Hydroxylated and methoxylated polybrominated diphenyl ethers and polybrominated dibenzo-p-dioxins in red alga and cyanobacteria living in the Baltic Sea. Chemosphere 2008, 72, 910-916. (52) Reddy, C. M.; Xu, L.; Eglinton, T. I.; Boon, J. P.; Faulkner, D. J. Radiocarbon content of synthetic and natural semi-volatile halogenated organic compounds. Environmental Pollution 2002, 120, 163-168. (53) Farid, S.; Dinnocenzo, J. P.; Merkel, P. B.; Young, R. H.; Shukla, D. Bimolecular Electron Transfers That Follow a Sandros−Boltzmann Dependence on Free Energy. Journal of the American Chemical Society 2011, 133, 4791-4801. (54) Farid, S.; Dinnocenzo, J. P.; Merkel, P. B.; Young, R. H.; Shukla, D.; Guirado, G. Reexamination of the Rehm–Weller Data Set Reveals Electron Transfer Quenching That Follows a Sandros–Boltzmann Dependence on Free Energy. Journal of the American Chemical Society 2011, 133, 11580-11587. (55) Sankoda, K.; Kuribayashi, T.; Nomiyama, K.; Shinohara, R. Occurrence and Source of Chlorinated Polycyclic Aromatic Hydrocarbons (Cl-PAHs) in Tidal Flats of the Ariake Bay, Japan. Environmental Science & Technology 2013, 47, 7037-7044.

Page 29: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 24  

Page 30: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  25  

Chapter 2

Photochemical formation of brominated

dioxins and other products of concern from

hydroxylated polybrominated diphenyl ethers

(OH-PBDEs) Paul R. Erickson, Matthew Grandbois, William A. Arnold and Kristopher

McNeill

Published in

Environmental science & Technology 2012, 46(15), 8174-8180

Page 31: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 26  

Abstract

The photochemical conversion of selected hydroxylated polybrominated

diphenyl ethers (OH-PBDEs) to dioxins and other products was investigated. OH-

PBDEs, which are both transformation products of polybrominated diphenyl ethers

and naturally occurring compounds, undergo direct photolysis to yield a number of

products that may have a higher toxicity than their parent. The compounds

investigated were 6-OH-PBDE 99, 6’-OH-PBDE 100, and 6’-OH-PBDE 118. Of

special interest was 6’-OH-PBDE 118, a potential transformation product of PBDE

153 that is capable of photochemically generating 2,3,7,8-tetrabromodibenzo-p-

dioxin, the most toxic brominated dioxin congener. Photolysis experiments were

conducted at two different pH values to assess the photochemical behavior of both the

phenol and phenolate form of the compounds. The percent conversion to dioxin and

other photoproducts was determined and the natural product, 6-OH-PBDE 99, was

found to have the highest conversion to dioxin (7%). The reaction quantum yields

ranged from 0.027 to 0.16 across all photolysis conditions. In addition, it is shown

that all three compounds are capable of photochemically generating other compounds

of concern, including brominated phenols and a dibenzofuran.

Page 32: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  27  

Introduction

Polybrominated diphenyl ethers (PBDEs) have been incorporated as flame-

retardants in a wide variety of consumer and industrial polymer products beginning in

the 1970s. The use of PBDEs as a non-covalently bound additive allows these

compounds to leach from the products into the environment. Consequently, PBDEs

are now ubiquitous in both urban and pristine environments alike.1,2 PBDEs entering

the environment have a number of potential fates, including biotic and abiotic

transformation, sorption to soils/sediment, and bioconcentration. One important class

of PBDE transformation products is comprised of the hydroxylated PBDEs (OH-

PBDEs).

Three main routes are thought to be responsible for the environmental

hydroxylation of PBDEs. The first is metabolic transformation in organisms exposed

to PBDEs. OH-PBDEs have been identified as metabolites in rats and mice3-5, and

OH-PBDE concentrations have been shown to have a positive correlation to PBDE

exposure in samples of human blood.6,7 Another possible path to hydroxylation is via

reaction with hydroxyl radical during atmospheric transport of PBDEs. Raff et al.

demonstrated that hydroxylation of PBDEs occurs in gas phase laboratory

experiments,8 and OH-PBDEs have been detected in rain and surface waters.9 PBDEs

may also become hydroxylated during the oxidative stages of wastewater treatment

processes,9 including activated sludge treatment and disinfection. The hydroxylation

of PBDEs increases their toxicity, because OH-PBDEs are stronger endocrine

disruptors10 and more closely resemble estrogenic compounds than PBDEs.11 OH-

PBDEs have been detected in a variety of freshwater12 and marine aquatic animals,

including salmon13, blue mussels14, dolphins15, and polar bears16. Further

complicating the picture, OH-PBDEs are not only transformation products of PBDEs,

Page 33: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 28  

but are also natural products found in marine systems. OH-PBDEs and closely related

compounds have been isolated from sponges17,18, tunicates19, and red algae14,20, and

radiocarbon experiments have further established their natural origin21,22.

In addition to the concern posed by OH-PBDEs themselves, congeners that

contain both a hydroxyl group ortho to the ether linkage and an ortho bromine on the

opposite ring are capable of undergoing a photochemical ring closure to form

polybrominated dibenzo-p-dioxins (PBDDs). This ring closure has been previously

observed for triclosan23,24, some of its chlorinated derivatives25, as well as OH-

PBDEs26,27. PBDDs are structurally analogous to their better-studied chlorinated

counterparts and share the same route of toxicity. Previous work has established that

PBDDs are at least as toxic as chlorinated dioxins and thus are a cause for concern28-

30. Alongside OH-PBDEs and other halogenated compounds found in marine

environments, there is some uncertainty as to the origin of PBDDs, which in some

places have been detected in increasing concentrations31. One possible source for

these dioxins may be photochemical formation from OH-PBDEs (Figure 1).

Page 34: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  29  

Figure 1. Transformation scheme illustrating possible routes from PBDEs to OH-

PBDEs (outside the scope of this study) and from OH-PBDEs to TBDDs.

The aim of this work was to investigate the photochemical degradation of

three different OH-PBDEs: 6-OH-PBDE 99 (1), 6’-OH-PBDE 100 (2) and 6’-OH-

PBDE 118 (3). Compound 1 was chosen because it is known to be a natural product,

and has been found in a variety of biota20,32. Compounds 2 and 3 were chosen because

they represent transformation products of two congeners present in the industrial

Penta mixture. Compound 3 is of particular interest since it is set up to undergo a

photochemical transformation to 2,3,7,8-TBDD, the most toxic dioxin congener. It is

also worth noting that 1 is a possible transformation product of BDE 99, which is also

one of the main components of the Penta mixture. Here we report that compounds 1-3

are each photochemically transformed under environmentally relevant conditions to

OBr

BrBr

Br

BrBr

PBDE 153

OOH

BrBr

Br

BrBr

6'-OH-PBDE 118 (3)

O

Br Br

2,3,7,8-TBDD

Brc or d

a: biotic hydroxylation b: + •OH, – H•; c: (photohydrolysis) h! + H2O, – HBr; d: + •OH, – Br•

O

Brh!– HBr

OBrOH

BrBr

Br

Br

6-OH-PBDE 99 (1)

O

O BrBrBr

Br1,2,4,8-TBDD

h!– HBr

Naturally occuringAnthropogenic

OOH

Br

Br

BrO

Br BrOh!

– HBrBrBr

PBDE 100 6'-OH-PBDE 100 (2) 1,3,7,9-TBDD

BrBr

O

Br

Br

BrBrBr

Focus of present study

OBr

BrBr

Br

Br

PBDE 99

a or b

a or b

Page 35: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 30  

polybrominated dioxins. In addition, we also report other photoproducts that are

concerning from a toxicity standpoint, including brominated phenols, a dibenzofuran,

and dihydroxybiphenyls.

Experimental

Chemicals

Compounds 1-3 were synthesized as described in Appendix 1. The

preparation of 6-OH PBDE 47 has been described previously.27 Pyridine (Py; 99.9%),

p-nitro anisole (PNA; 97%), NaOH, HCl, acetic acid, and sodium acetate were all

purchased from Sigma Aldrich. Sodium tetraborate decahydrate (Na2B4O7·10H2O)

was purchased from Merck. The standards used were 1,2,4,8/1,2,4,7-

tetrabromodibenzo-p-dioxin (10 µg/mL in toluene, AccuStandard), 2,3,7,8-

tetrabromodibenzo-p-dioxin (50 µg/mL in toluene, Wellington Labs), 2,4-

dibromophenol (neat, Supelco), 2,4,5-tribromophenol (1 mg/mL in isopropanol,

Chiron AS), and 2,4,6-tribromophenol (neat, Sigma Aldrich). All compounds were

used as received with the exception of p-nitro anisole (PNA), which was first

recrystallized from n-hexane. Ultrapure water (18 MΩ·cm) was obtained from a

Barnstead Nanopure Diamond system. All solvents were of chromatography grade.

pKa value determinations

pKa values were measured by spectrophotometric titration. Because the water

solubility of these compounds is very low when they are in the protonated form,

titrations had to be performed in methanol:water mixtures. Solutions of 1-3 (15 mL,

approx. 50 µM) in various methanol/water mixtures were brought to a pH value

between 10 and 11 with NaOH. These solutions were then titrated to a pH value near

Page 36: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  31  

4 with a maximum of 200 µL of 0.1 M HCl, so that the change in OH-PBDE

concentration was negligible. Absorbance measurements were taken with a Cary 100

Bio (Varian) UV/Visible spectrophotometer in 1.00 cm quartz cuvettes. Plots of the

absorbance at ~ 310 nm vs pH were fit using a nonlinear regression (Kaleidograph,

version 4.04, Synergy Software) to determine the pKa values.

Photolysis experiments

Samples of 1-3 in 1 mM pH 4 acetate buffer with 30% methanol or 1 mM pH

10 borate buffer with an initial concentration of 10 µM were placed in a merry-go-

round sample holder of a Rayonet photoreactor containing two 300 nm bulbs

(spectrum from within a test tube shown in Figure 2). The irradiation intensity in the

range of 280-400 nm was about 40% of natural sunlight. For all photolysis

experiments, triplicate 10 mL samples were irradiated in borosilicate glass test tubes.

At set intervals, 150 µL aliquots were taken for analysis by either high-pressure liquid

chromatography (HPLC) for kinetics and photoproduct quantification or liquid

chromatography-mass spectrometery (LC-MS) for photoproduct identification. Dark

controls showed no loss of parent compound over 2h irradiation time. To determine

accurate polychromatic apparent reaction quantum yield values (Φr), a PNA-Py

chemical actinometer system33 was irradiated alongside the samples in borosilicate

glass test tubes. Φr values for 1-3 were calculated according to equation 1.

(1)

Φ r =krka

εa,λIλ∑εa,λIλ∑

Φa

Page 37: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 32  

First-order rate constants of degradation (kdeg) were compared to that of a PNA-Py

chemical actinometer with a known Φa. Molar absorptivities (ε) as a function of

wavelength were measured for each compound at pH 4 and 10, and the lamp spectrum

was measured from inside a borosilicate test tube using a fiber optic coupled to a

calibrated radiometer (Ocean Optics Jaz).

The percent conversion is defined as the rate constant of formation of a given

photoproduct (kp) divided by the overall rate constant of degradation (kdeg) of the

parent compound multiplied by 100. It is not possible to measure kp directly from the

growth of product, because product appearance is kinetically bound to the rate

constant of starting material loss kdeg. For a photostable product, the ratio kp/kdeg is

equal to the final product concentration [P]∞ divided by the starting concentration of

the parent compound [A]0. Dioxin percent conversion values were calculated in this

manner because of their negligible photodegradation over the course of the

experiments. For the bromophenols, which were photolabile, the kinetic profiles were

fit to equation 2. The term f[A]0 refers to the fraction of A that goes to form

photoproduct P (i.e., the percent conversion or kp/kdeg), and the percent conversion

values were calculated based on the f[A]0 fit parameter.

[P]=kdeg fAokp − kdeg

(e−kdeg ⋅t − e−kp ⋅t ) (2)

Page 38: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  33  

Photoproduct identification and quantification

When photoproducts of 1-3 were commercially available, these compounds

were used to create calibration curves for the HPLC and LC-MS (except for dioxins,

which were only quantified by HPLC). All calibration curves used for quantification

had an R2 of 0.999 or better. Only photoproducts for which an authentic standard was

available were quantified and a percent conversion reported, with the exception of

1,3,7,9-TBDD, which was quantified using the calibration curve for 2,3,7,8-TBDD.

This was deemed reasonable because the slopes of the calibration curves for the other

two tetrabrominated congener standards, 1,2,4,8- and 2,3,7,8-TBDD were within 2%

of one another. The identity of 1,3,7,9-TBDD was inferred from the precursor

structure reaction scheme in Figure 1 and on the basis of a photoproduct with an

elution time and UV-vis that were similar to the two standards that were available.

Photoproducts for which no authentic standards were available were identified and

classified only by their general structure using exact mass measurements and relative

retention time data. Products observed by LC-MS alone could not be quantified,

because the MS instrument response strongly depends on the ionization efficiency of

the compound.

HPLC analysis

Compounds 1-3 and actinometer samples were analyzed using a Dionex P680

HPLC with a PDA-100 photodiode array UV-vis detector. Injection volumes of 100

mL were used for all samples. The column was a Supelco Discovery C16 RP amide

(15 cm × 4.6 mm, 5 µm dia. particles). For analysis of 1-3, the mobile phase was

85:13.5:1.5 methanol:pH 3 phosphate buffer:acetonitrile at a flow rate of 1 mL/min.

All compounds were quantified at a wavelength of 210 nm except for the dioxins,

which were quantified at 230 nm. Sample run times were 35 minutes. For the

Page 39: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 34  

actinometer, the mobile phase was 65:31.5:3.5 methanol:pH 3 phosphate

buffer:acetonitrile at a flow rate of 1 mL/min. PNA was quantified at a wavelength of

314 nm. Sample run times were 7 minutes.

LC-MS analysis

Photoproducts of 1-3 were identified using a Waters NanoAquity UPLC

interfaced to a Thermo Exactive Orbitrap high-resolution mass spectrometer.

Injections (1 µL) were made onto a Waters Atlantis dC18 Nano Ease column (15 cm

× 300 µm, 3 µm dia. particles) held at 40° C with a flow rate of 17 µL min-1 and a

mobile phase composition of 60:40 acetonitrile:water. All samples were analyzed by

electrospray ionization MS in negative mode and were scanned from 100-600 m/z.

Sample run times were 15 min.

Results and Discussion

pKa values

The pKa values for the phenolic groups of 1-3 were 5.81, 8.39, and 7.39,

respectively (Table 1). The pKa value for 1, which contains three bromo substituents

on the phenol ring, is significantly lower than for 2 and 3, which each only contain

two bromo substitutents on the phenol ring. A series of titrations for compounds 1 and

2 in different methanol:water compositions demonstrated linear relationships with the

solvent composition that could be used to extrapolate to the pure aqueous pKa values.

Compound 3 was titrated only in 50:50 methanol:water because of material

limitations. A reasonable prediction of the aqueous pKa value was made for 3 using

the common slope from the extrapolation plots for 1 and 2. The titration plots and

extrapolations are provided in Appendix 1.

Page 40: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  35  

Table 1. pKa values, reaction quantum yields, and quantified photoproducts

Direct photolysis

The loss of compounds 1-3 during irradiation followed first-order kinetics

under all experimental conditions tested. The starting concentration for all photolysis

samples was low enough that self-screening was unimportant, making it unnecessary

to correct kinetic data used to calculate Φr values. Figure 2 shows the absorption

spectra and degradation plots for 1-3. In the phenol form, all of the compounds have a

broad absorbance band with a maximum near 290 nm, tailing off to approximately

zero absorbance at around 310 nm. In the phenolate form, the absorbance maximum

red-shifts for each compound differently, but they all display a maximum between

300 and 310 nm and a tailing off to near zero absorbance around 340 nm. All

Parent compound pKa

Debrominationc(% conversion)

C-O cleavagec(% conversion)

Dioxinc(% conversion)

OBrOH

BrBr

Br

Br

6-OH-BDE 99 (1)

OOH

Br

Br

Br

6'-OH-BDE 100 (2)

BrBr

OOH

BrBr

Br

BrBr

6'-OH-BDE 118 (3)

5.81a 0.0270.031

8.39a 0.0630.16

0.0570.137.39b

O

O BrBrBr

Br1,2,4,8-TBDD

(7%)

O

Br BrO

BrBr

1,3,7,9-TBDD(1.1%)

O

Br Br

2,3,7,8-TBDD(0.5%)

Br

O

Br

!pH 4pH 10

OBrOH

BrBr Br

6-OH-BDE 47(<1%)

OHBr

Br

OHBr

Br

OHBr

Br

Br

Br

2,4-dibromo-phenol(8%)

2,4,6-tribromo-phenol(11%)

2,4,5-tribromo-phenol(13%)

a: Estimated pure aqueous pKa value determined from a series of titrations of solutions with different methanol:water compositions. pKa value given is the extrapolated value. b: Estimated pKa value using the slope from the other compounds extrapolation (both lines have the same slope) and the pKa value determined in a 50% methanol solution. c: Quantification is reported for pH 10 experiments. d: Photoproducts for which no standard was available were not quantified.

OBrOH

BrBr Br

6-OH-BDE 47(<1%)

n.q.d

Page 41: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 36  

compounds degraded faster under basic conditions due mostly to the greater light

absorption of the phenolate species.

Figure 2. The absorbance spectra for the phenol and phenolate forms of compounds 1-3 along with the lamp spectrum are shown in panels (A) and (B), respectively. Panels (C) and (D) show kinetic traces for the photodegradation of 1-3 at pH 4 and 10, respectively.

For the phenol forms of 1-3, Φr values ranged from 0.027 to 0.063, while the

values for the phenolate forms ranged between 0.031 and 0.16 (Table 1). Our results

are in good agreement with earlier work with halogenated phenoxyphenols.23-27 Under

both acidic and basic conditions, 2 and 3 had noticeably larger Φr values than 1. One

possible contributing factor to this observation is the bromine in the 3 position of 1,

ortho to the hydroxyl group. While the origin of this “ortho effect” on quantum yield

0

1000

2000

3000

4000

5000

6000

7000

0

0.2

0.4

0.6

0.8

1

280 290 300 310 320 330 340 350

Mol

ar a

bsor

btiv

ity (M

-1 c

m-1

)

Lamp spectral intensity (arb. units)

Wavelength (nm)

1

2

3 A.

0

1000

2000

3000

4000

5000

6000

7000

0

0.2

0.4

0.6

0.8

1

280 290 300 310 320 330 340 350

Mol

ar a

bsor

btiv

ity (M

-1 c

m-1

)

Lamp spectral intensity (arb. units)

Wavelength (nm)

1

2

3

B.

-2

-1.5

-1

-0.5

0

0 10 20 30 40

ln(C

/Co)

Time (min)

2

3

1

1= 6-OH-PBDE 992= 2'-OH-PBDE 1003= 6'-OH-PBDE 118

C.

-3

-2.5

-2

-1.5

-1

-0.5

0

0 10 20 30 40

ln(C

/Co)

Time (min)

123

D.

Page 42: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  37  

is not understood, it is noteworthy, particularly in light of our previous work which

showed the same effect for chlorinated triclosan derivatives.25

Brominated Dioxins

Dioxin formation was observed for 1-3 under all conditions. The percent

conversion to dioxin for the compounds at pH 10 was found to be 7, 1.1, and 0.5%. In

all cases, the dioxin photoproducts were relatively photostable under the experimental

conditions, and were seen to photodegrade on the order of 100 times slower than the

parent compounds during our experiments. The factors that control the wide range in

dioxin conversion yields are not understood. Nevertheless, it is a happy circumstance

that compound 3, which forms the most toxic dioxin congener, 2,3,7,8-TBDD, had the

lowest dioxin yield of 0.5%. By contrast, compound 1, the natural product, was found

to have an order of magnitude higher dioxin conversion of 7%.

Bromophenols

Compounds 1-3 all generated a bromophenol photoproduct from the cleavage

of the ether linkage where the bridging oxygen atom remains with the non-phenolic

ring. Interestingly, these were the only C-O cleavage products observed above trace

levels in any of our experiments. The percent conversion values ranged from 8-13%.

The amount of bromophenol photoproduct is not controlled by steric crowding around

the ether linkage. Indeed, compound 3, the least sterically crowded congener, showed

the highest bromophenol conversion. Noting that the higher conversion yields are

associated with a greater number of bromo substitutents on the departing phenol, the

bromophenol production may instead be a function of the resulting bromophenolate’s

leaving group ability. A study with a wider range of structural analogs will be needed

to more rigorously test this hypothesis.

Page 43: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 38  

Debromination and di-OH-PBB formation

Hydrodebromination is one of the predominant degradation pathways for

PBDEs34,35, therefore we expected to observe a number of transformation products

corresponding to reductive debromination. Instead, we observed only small amounts

of debromination products. This conclusion was based on two sets of observations.

First, we quantified the production of 6-OH-PBDE 47, the only debromination

product standard that we had available. For 1 to form 6-OH-PBDE 47 it must lose the

bromine ortho to the hydroxy group, while 2 must lose one of its two bromo

substituents ortho to the ether linkage on the non-hydroxylated ring. In both cases, 6-

OH-PBDE 47 is formed in less than 1% yield. Second, we examined LC-MS ion

chromatograms for each of the hydrodebromination product masses. We observed a

group of photoproducts with the same formula (and thus exact mass) as the reductive

debromination products, which are believed to belong to another compound class.

Based on their poor retention on the HPLC column (ca. 3 min), which is similar to

bromophenols, and literature precedent for diphenyl ether to hydroxy biphenyl

photoconversion34, we tentatively assign these early eluting products as

dihydroxylated polybrominated biphenyls (di-OH-PBBs). The analogous

photochemical formation of OH-PBBs from a PBDE has been reported from the

photolysis of BDE-153 in 20% yield.34 Due to a lack of authentic standards and an

inability to collect enough material for structural identification by NMR, a firm

assignment could not be made. Tentatively assigned debromination products, which

have retention times near the parent OH-PBDEs, and di-OH-PBBs were observed in

varying amounts for 1-3, but all compounds produced at least one of each.

Page 44: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  39  

Dihydroxy-PBDEs and hydroxydibenzofuran

The last two types of photoproducts that were identified were products arising

from the replacement of bromine with a hydroxyl group to form di-OH-PBDEs, and

products that had formally lost HBr to form a hydroxylated polybrominated

dibenzofuran (OH-PBDF).

The di-OH-PBDEs are isobaric with tri-OH-PBBs and thus difficult to

distinguish by MS. Furthermore, both compound types are expected to elute early in

reverse phase LC. Indeed, the peaks assigned as di-OH-PBDEs had early retention

times in the LC-MS chromatogram, similar to bromophenols. We favor the

assignment of di-OH-PBDEs (versus tri-OH-PBBs) due to their kinetic behavior.

Specifically, the plot of their appearance vs. time (figure S8) does not show any lag in

initial production, as would be expected for a second-generation transformation

product (i.e., arising from hydroxy-debromination of di-OH-PBBs).

Finally, we observed the formation of a photoproduct with a retention time

similar to that of the dioxin products during HPLC analysis, and with a mass

corresponding to the loss of HBr for compound 2. We propose that this product is a

hydroxy-substituted polybrominated dibenzofuran (OH-PBDF). The large peak in the

HPLC chromatogram suggests that a significant amount of the dibenzofuran

photoproduct was formed. It is also worth noting that the photochemical formation of

hydroxylated polychlorinated dibenzofurans has been previously observed from

triclosan and its chlorinated derivatives.36,37

Because no authentic standards for OH-PBDFs are commercially available, we

synthesized a tetrabrominated OH-PBDF to judge whether the UV-vis absorbance

Page 45: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 40  

spectrum and HPLC retention time was similar to that of the suspected OH-PBDF.

The choice of the synthetic standard was constrained by its synthetic accessibility.

While not a match to our proposed product, the standard was helpful due to the

similarity of dibenzofuran retention times and absorbance spectra for congeners of the

same degree of halogenation. The details of the synthesis can be found in Appendix

1. Our synthesized OH-PBDF had an HPLC retention time of 23.5 min and the

suspected OH-PBDF produced by 2 had a retention time of 24.3 min, respectively

(Figure 3). A comparison of their UV-vis spectra shows three similar spectral

features and an absorbance tailing to zero near 330 nm. Based on this comparison and

the exact mass match, we favor the assignment of the unknown as a tetrabrominated

OH-PBDF, but cannot definitively assign the structure. Based on our proposed

transformation mechanism (discussed below), we propose that 1,2,6,8-tetrabromo-4-

hydroxydibenzofuran is the specific congener formed from 2.

At first glance, the formation of OH-PBDF from 2 seems unlikely, because it

does not have any hydrogen atoms ortho to the ether linkage to eliminate during

cyclization. Our hypothesis is that furan formation proceeds via a photochemical

electrocyclic ring closure to form the ring system skeleton,38 followed by a 1,3-

migration of a bromine atom from the non-phenolic ring to the phenolic ring (Figure

S9). Subsequent elimination and tautomerization would yield the proposed product

(Figures 3 and S9).

Page 46: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  41  

Figure 3. (Top panels) The HPLC-UV (diode array) chromatogram of the product mixture resulting from the photolysis of 2 and the UV absorbance spectrum corresponding to the indicated peak. The peak eluting at 11.7 min is of the starting compound 2. (Bottom panels) The HPLC-UV chromatogram is of the synthesized OH-TBDF and the UV absorbance spectrum of the indicated peak. See text for assignment of the dibenzofuran structures.

Environmental significance

The presence of OH-PBDEs in the environment is now well established, and

while the largest share of OH-PBDEs comes from natural sources, the possibility of

anthropogenic sources remain. Our results show that 1-3 all photodegrade to generate

transformation products that are themselves priority pollutants. While bromophenol

production was seen for all compounds, the anthropogenically derived compounds, 2

and 3, produced the largest amounts. In addition to generating bromophenols, the

previously unreported formation of di-OH-PBBs and OH-PBDFs from OH-PBDEs

also could be of concern. Compound 1 is a known natural product, and has been

0

50

100

150

200

250

300

0 5 10 15 20 25 30 35

Tetrabromo OH-dibenzofuran

Abs

orba

nce

( arb

. uni

ts)

Time (min)

0

500

1000

1500

2000

2500

200 250 300 350 400

Tetrabromo OH-dibenzofuran

Abso

rban

ce (a

rb. u

nits

)

Wavelength (nm)

0

50

100

150

200

200 250 300 350 400

Unknown photoproduct

Abso

rban

ce (a

rb. u

nits

)

Wavelength (nm)

OOH

Br

BrBr

OOH

Br

Br

Br

Proposed structure

0

10

20

30

40

50

60

0 5 10 15 20 25 30 35

OH-PBDE 100 photolysis T= 2minA

bsor

banc

e ( a

rb. u

nits

)

Time (min)

OOH

Br

Br

BrBrBr Br

Br

Page 47: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 42  

isolated from biotic sources in the Baltic Sea.14 Our results show that 1 is readily

capable of transforming into a TBDD under environmental conditions, and may be a

major source of natural tetrabrominated dioxin in the Baltic Sea, a possibility that has

been previously discussed.39 While 1 produced more dioxin than 2 and 3, the toxicity

of the resulting dioxin is likely much higher for 2 and certainly for 3. The scarcity of

toxicity data and lack of accepted toxic equivalent factors (TEFs) for the dioxins

formed by 1 and 2 complicates the comparison of their relative dioxin toxicity. If we

conservatively assume a TEF of 0.001 (the actual TEF is likely much lower) for

1,2,4,8-TBDD, then at the dioxin production rates from our experiments, compound 3

will produce 70 times the toxicity as the natural product 1 from the same starting

concentrations.

Our work confirms that OH-PBDEs of both natural and anthropogenic origin

are capable of being converted photochemically to a variety of other brominated

compounds, including brominated dioxins. While the actual production rates for the

anthropogenic OH-PBDEs remain unknown, this work supports the hypothesis that

these compounds, whether natural or anthropogenic, are a source of brominated

dioxins in the environment. Further work will be required to better constrain any

estimates about the share of dioxin toxicity attributable to the various OH-PBDE

sources.

Acknowledgements

We thank Prof. P.J. Alaimo (Seattle Univ.) for the helpful discussions. We

gratefully acknowledge support of the U.S. National Science Foundation (CBET-

0967163).

Page 48: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  43  

References

(1) de Wit, C. A. An overview of brominated flame retardants in the environment. Chemosphere 2002, 46, 583-624.

(2) Hale, R. C.; Alaee, M.; Manchester-Neesvig, J. B.; Stapleton, H. M.; Ikonomou, M. G. Polybrominated diphenyl ether flame retardants in the North American environment. Environment International 2003, 29, 771-779.

(3) Malmberg, T.; Athanasiadou, M.; Marsh, G.; Brandt, I.; Bergman, Å. Identification of Hydroxylated Polybrominated Diphenyl Ether Metabolites in Blood Plasma from Polybrominated Diphenyl Ether Exposed Rats. Environmental Science & Technology 2005, 39, 5342-5348.

(4) Marsh, G.; Athanasiadou, M.; Athanassiadis, I.; Sandholm, A. Identification of hydroxylated metabolites in 2,2′,4,4′-tetrabromodiphenyl ether exposed rats. Chemosphere 2006, 63, 690-697.

(5) Örn, U.; Klasson-Wehler, E. Metabolism of 2,2, 4,4-tetrabromodiphenyl ether in rat and mouse. Xenobiotica 1998, 28, 199-211.

(6) Athanasiadou, M.; Cuadra, S. N.; Marsh, G.; Bergman, Å.; Jakobsson, K. Polybrominated Diphenyl Ethers (PBDEs) and Bioaccumulative Hydroxylated PBDE Metabolites in Young Humans from Managua, Nicaragua. Environ Health Perspect 2007, 116.

(7) Stapleton, H. M.; Eagle, S.; Anthopolos, R.; Wolkin, A.; Miranda, M. L. Associations between Polybrominated Diphenyl Ether (PBDE) Flame Retardants, Phenolic Metabolites, and Thyroid Hormones during Pregnancy. Environ Health Perspect 2011, 119.

(8) Raff, J. D.; Hites, R. A. Gas-Phase Reactions of Brominated Diphenyl Ethers with OH Radicals. The Journal of Physical Chemistry A 2006, 110, 10783-10792.

(9) Ueno, D.; Darling, C.; Alaee, M.; Pacepavicius, G.; Teixeira, C.; Campbell, L.; Letcher, R. J.; Bergman, Å.; Marsh, G.; Muir, D. Hydroxylated Polybrominated Diphenyl Ethers (OH-PBDEs) in the Abiotic Environment: Surface Water and Precipitation from Ontario, Canada. Environmental Science & Technology 2008, 42, 1657-1664.

(10) Cantón, R. F.; Scholten, D. E. A.; Marsh, G.; de Jong, P. C.; van den Berg, M. Inhibition of human placental aromatase activity by hydroxylated polybrominated diphenyl ethers (OH-PBDEs). Toxicology and Applied Pharmacology 2008, 227, 68-75.

(11) Mercado-Feliciano, M.; Bigsby, R. M. Hydroxylated Metabolites of the Polybrominated Diphenyl Ether Mixture DE-71 Are Weak Estrogen Receptor-α Ligands. Environ Health Perspect 2008, 116.

(12) Valters, K.; Li, H.; Alaee, M.; D'Sa, I.; Marsh, G.; Bergman, Å.; Letcher, R. J. Polybrominated Diphenyl Ethers and Hydroxylated and Methoxylated Brominated and Chlorinated Analogues in the Plasma of Fish from the Detroit River. Environmental Science & Technology 2005, 39, 5612-5619.

(13) Marsh, G.; Athanasiadou, M.; Bergman, Å.; Asplund, L. Identification of Hydroxylated and Methoxylated Polybrominated Diphenyl Ethers in Baltic Sea Salmon (Salmo salar) Blood. Environmental Science & Technology 2004, 38, 10-18.

(14) Malmvärn, A.; Marsh, G.; Kautsky, L.; Athanasiadou, M.; Bergman, Å.; Asplund, L. Hydroxylated and Methoxylated Brominated Diphenyl Ethers in the Red Algae Ceramium tenuicorne and Blue Mussels from the Baltic Sea. Environmental Science & Technology 2005, 39, 2990-2997.

Page 49: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 44  

(15) Houde, M.; Pacepavicius, G.; Darling, C.; Fair, P. A.; Alaee, M.; Bossart, G.

D.; Solomon, K. R.; Letcher, R. J.; Bergman, Å.; Marsh, G.; Muir, D. C. G. Polybrominated diphenyl ethers and their hydroxylated analogs in plasma of bottlenose dolphins (Tursiops truncatus) from the United States east coast. Environmental Toxicology and Chemistry 2009, 28, 2061-2068.

(16) Verreault, J.; Gabrielsen, G. W.; Chu, S.; Muir, D. C. G.; Andersen, M.; Hamaed, A.; Letcher, R. J. Flame Retardants and Methoxylated and Hydroxylated Polybrominated Diphenyl Ethers in Two Norwegian Arctic Top Predators:   Glaucous Gulls and Polar Bears. Environmental Science & Technology 2005, 39, 6021-6028.

(17) Handayani, D.; Edrada, R. A.; Proksch, P.; Wray, V.; Witte, L.; Van Soest, R. W. M.; Kunzmann, A.; Soedarsono Four New Bioactive Polybrominated Diphenyl Ethers of the Sponge Dysidea herbacea from West Sumatra, Indonesia. Journal of Natural Products 1997, 60, 1313-1316.

(18) Carté, B.; Faulkner, D. J. Polybrominated diphenyl ethers from Dysidea herbacea, Dysidea chlorea and Phyllospongia foliascens. Tetrahedron 1981, 37, 2335-2339.

(19) Schumacher, R. W.; Davidson, B. S. Didemnolines A-D, new N9-substituted β-carbolines from the marine ascidian Didemnum sp. Tetrahedron 1995, 51, 10125-10130.

(20) Malmvärn, A.; Zebühr, Y.; Kautsky, L.; Bergman, Å.; Asplund, L. Hydroxylated and methoxylated polybrominated diphenyl ethers and polybrominated dibenzo-p-dioxins in red alga and cyanobacteria living in the Baltic Sea. Chemosphere 2008, 72, 910-916.

(21) Reddy, C. M.; Xu, L.; Eglinton, T. I.; Boon, J. P.; Faulkner, D. J. Radiocarbon content of synthetic and natural semi-volatile halogenated organic compounds. Environmental Pollution 2002, 120, 163-168.

(22) Teuten, E. L.; Xu, L.; Reddy, C. M. Two Abundant Bioaccumulated Halogenated Compounds Are Natural Products. Science 2005, 307, 917-920.

(23) Latch, D. E.; Packer, J. L.; Arnold, W. A.; McNeill, K. Photochemical conversion of triclosan to 2,8-dichlorodibenzo-p-dioxin in aqueous solution. Journal of Photochemistry and Photobiology A: Chemistry 2003, 158, 63-66.

(24) Latch, D. E.; Packer, J. L.; Stender, B. L.; VanOverbeke, J.; Arnold, W. A.; McNeill, K. Aqueous photochemistry of triclosan: Formation of 2,4-dichlorophenol, 2,8-dichlorodibenzo-p-dioxin, and oligomerization products. Environmental Toxicology and Chemistry 2005, 24, 517-525.

(25) Buth, J. M.; Grandbois, M.; Vikesland, P. J.; McNeill, K.; Arnold, W. A. Aquatic photochemistry of chlorinated triclosan derivatives: Potential source of polychlorodibenzo-P-dioxins. Environmental Toxicology and Chemistry 2009, 28, 2555-2563.

(26) Bastos, P. M.; Eriksson, J.; Bergman, Å. Photochemical decomposition of dissolved hydroxylated polybrominated diphenyl ethers under various aqueous conditions. Chemosphere 2009, 77, 791-797.

(27) Steen, P. O.; Grandbois, M.; McNeill, K.; Arnold, W. A. Photochemical Formation of Halogenated Dioxins from Hydroxylated Polybrominated Diphenyl Ethers (OH-PBDEs) and Chlorinated Derivatives (OH-PBCDEs). Environmental Science & Technology 2009, 43, 4405-4411.

(28) Ao, K.; Suzuki, T.; Murai, H.; Matsumoto, M.; Nagai, H.; Miyamoto, Y.; Tohyama, C.; Nohara, K. Comparison of immunotoxicity among tetrachloro-,

Page 50: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  45  

pentachloro-, tetrabromo- and pentabromo-dibenzo-p-dioxins in mice. Toxicology 2009, 256, 25-31.

(29) Birnbaum, L. S.; Staskal, D. F.; Diliberto, J. J. Health effects of polybrominated dibenzo-p-dioxins (PBDDs) and dibenzofurans (PBDFs). Environment International 2003, 29, 855-860.

(30) Van den Berg, M.; Birnbaum, L. S.; Denison, M.; De Vito, M.; Farland, W.; Feeley, M.; Fiedler, H.; Hakansson, H.; Hanberg, A.; Haws, L.; Rose, M.; Safe, S.; Schrenk, D.; Tohyama, C.; Tritscher, A.; Tuomisto, J.; Tysklind, M.; Walker, N.; Peterson, R. E. The 2005 World Health Organization Reevaluation of Human and Mammalian Toxic Equivalency Factors for Dioxins and Dioxin-Like Compounds. Toxicological Sciences 2006, 93, 223-241.

(31) Haglund, P.; Malmvärn, A.; Bergek, S.; Bignert, A.; Kautsky, L.; Nakano, T.; Wiberg, K.; Asplund, L. Brominated Dibenzo-p-Dioxins:   A New Class of Marine Toxins? Environmental Science & Technology 2007, 41, 3069-3074.

(32) Löfstrand, K.; Liu, X.; Lindqvist, D.; Jensen, S.; Asplund, L. Seasonal variations of hydroxylated and methoxylated brominated diphenyl ethers in blue mussels from the Baltic Sea. Chemosphere 2011, 84, 527-532.

(33) Leifer, A. The Kinetics of Environmental Aquatic Photochemistry; American Chemical Society: Washington, DC, 1988.

(34) Rayne, S.; Wan, P.; Ikonomou, M. Photochemistry of a major commercial polybrominated diphenyl ether flame retardant congener: 2,2′,4,4′,5,5′-Hexabromodiphenyl ether (BDE153). Environment International 2006, 32, 575-585.

(35) Söderström, G.; Sellström, U.; de Wit, C. A.; Tysklind, M. Photolytic Debromination of Decabromodiphenyl Ether (BDE 209). Environmental Science & Technology 2003, 38, 127-132.

(36) Kanetoshi, A.; Ogawa, H.; Katsura, E.; Kaneshima, H. Chlorination of irgasan DP300 and formation of dioxins from its chlorinated derivatives. Journal of Chromatography A 1987, 389, 139-153.

(37) Sanchez-Prado, L.; Llompart, M.; Lores, M.; García-Jares, C.; Bayona, J. M.; Cela, R. Monitoring the photochemical degradation of triclosan in wastewater by UV light and sunlight using solid-phase microextraction. Chemosphere 2006, 65, 1338-1347.

(38) Pollard, R.; Wan, P. Synthetic applications of diaryl ether photochemistry. A review. Organic Preparations and Procedures International 1993, 25, 1-14.

(39) Haglund, P. On the identity and formation routes of environmentally abundant tri- and tetrabromodibenzo-p-dioxins. Chemosphere 2009, 78, 724-730.

Page 51: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 46  

Appendix to chapter 2 pKa determinations of compounds 1-3

Figure A1- Plot of observed pKa vs. % methanol for all titration experiments. For 6-OH-PBDE 99 (1) and 6’-OH-PBDE 100 (2) the extrapolation lines used to find the pure water pKa are shown. The pKa for 6’-OH-PBDE 118 (3) in 50% methanol is also shown.

Figure A2- Titration curve for 6’-OH-PBDE 118

5

6

7

8

9

10

0 10 20 30 40 50 60 70

y = 8.3927 + 0.01794x R= 0.96433

y = 5.8133 + 0.017x R= 0.9909

Obs

erve

d pK

a

% Methanol

2

3

1

0

0.05

0.1

0.15

0.2

0.25

0.3

6 7 8 9 10 11 12

6'-OH-PBDE 118 (50% MeOH)

Abso

rban

ce a

t 310

nm

pH

y = (10^(-m0)/(10^(-m1)+10^(...ErrorValue

0.0243898.264m1 0.00248670.036926m2 0.0018470.26553m3

NA0.00033089ChisqNA0.99889R

Page 52: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  47  

Figure A3- Titration curves for 6-OH-PBDE 99

0

0.05

0.1

0.15

3 4 5 6 7 8 9 10 11

6-OH-PBDE 99 (20% MeOH)

Abso

rban

ce a

t 309

nm

pH

y = (10^(-m0)/(10^(-m1)+10^(...ErrorValue

0.0646946.1371m1 0.00393320.0050693m2 0.00175560.13817m3

NA7.7093e-5ChisqNA0.99744R

0

0.02

0.04

0.06

0.08

0.1

0.12

0.14

3 4 5 6 7 8 9 10 11

6-OH-PBDE 99 (30% MeOH)

Abso

rban

ce a

t 309

nm

pH

y = (10^(-m0)/(10^(-m1)+10^(...ErrorValue

0.0290386.3494m1 0.00132910.008691m2 0.00104450.1359m3

NA7.3122e-5ChisqNA0.9991R

0

0.05

0.1

0.15

3 4 5 6 7 8 9 10 11

6-OH-PBDE 99 (40% MeOH)

Abso

rban

ce a

t 309

nm

pH

y = (10^(-m0)/(10^(-m1)+10^(...ErrorValue

0.0376076.4802m1 0.00167030.010706m2 0.00143710.13652m3

NA7.303e-5ChisqNA0.99886R

Page 53: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 48  

Figure A4- Titration curves for 6’-OH-PBDE 100

0

0.02

0.04

0.06

0.08

0.1

0.12

4 5 6 7 8 9 10 11

6'-OH-PBDE 100 (30% MeOH)

Abso

rban

ce a

t 310

nm

pH

y = (10^(-m0)/(10^(-m1)+10^(...ErrorValue

0.0656628.9136m1 0.00271690.020109m2 0.00179870.11216m3

NA0.00027083ChisqNA0.99346R

0

0.02

0.04

0.06

0.08

0.1

5 6 7 8 9 10 11

6'-OH-PBDE 100 (40% MeOH)

Abso

rban

ce a

t 310

nm

pH

y = (10^(-m0)/(10^(-m1)+10^(...ErrorValue

0.0632389.0962m1 0.002360.0084661m2

0.0021160.10309m3 NA0.00014744ChisqNA0.99569R

0

0.02

0.04

0.06

0.08

0.1

5 6 7 8 9 10 11

6'-OH-PBDE 100 (50% MeOH)

Abso

rban

ce a

t 310

nm

pH

y = (10^(-m0)/(10^(-m1)+10^(...ErrorValue

0.0447119.3813m1 0.00135780.0077988m2 0.00178930.10091m3

NA5.7293e-5ChisqNA0.99808R

0

0.02

0.04

0.06

0.08

0.1

0.12

3 4 5 6 7 8 9 10 11

6'-OH-PBDE 100 (60% MeOH)

Abso

rban

ce a

t 310

nm

pH

y = (10^(-m0)/(10^(-m1)+10^(...ErrorValue

0.0607329.413m1 0.00231570.0058497m2 0.00403970.12106m3

NA0.00010461ChisqNA0.99628R

Page 54: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  49  

Selected extracted ion chromatograms

Figure A5- Extracted ion mass chromatograms for OH-PBDE 99. All ions were observed as [M-H]-.

0 2 4 6 8 100

5 104

1 105

1.5 105

2 105

2.5 105

3 105

Inte

nsity

Elution time (min)

OBrOH

BrBr

Br

Br

0 2 4 6 8 100

5 1041 105

1.5 1052 105

2.5 1053 105

3.5 105

Inte

nsity

Elution time (min)

OH

Br

Br

0 2 4 6 8 100

2 106

4 106

6 106

8 106

1 107In

tens

ity

Elution time (min)

0 2 4 6 8 100

5 1041 105

1.5 1052 105

2.5 1053 105

3.5 1054 105

Inte

nsity

Elution time (min)

0 2 4 6 8 100

2000

4000

6000

8000

1 104

1.2 104

Inte

nsity

Elution time (min)

OH

Br4

OOH

Br4

OH

OOH

Br4

OOH

Br4

OH

m/z = 578.608

OBrOH

BrBr Br

m/z = 500.698

m/z = 250.853

m/z = 516.693

m/z = 498.682

Page 55: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 50  

Figure A6- Extracted ion mass chromatograms for OH-PBDE 100. All ions were observed as [M-H]-.

0 2 4 6 8 100

5 104

1 105

1.5 105

2 105

2.5 105

3 105

Inte

nsity

Elution time (min)

0 2 4 6 8 100

2 1054 1056 1058 1051 106

1.2 1061.4 1061.6 106

Inte

nsity

Elution time (min)

OBrOH

BrBr

OH

Br

Br

OH

Br4

OOH

Br4

OH

OOH

Br4

OOH

Br4

OH

0 2 4 6 8 100

5 1041 105

1.5 1052 105

2.5 1053 105

3.5 105

Inte

nsity

Elution time (min)

0 2 4 6 8 100

1 104

2 104

3 104

4 104

5 104

Inte

nsity

Elution time (min)

0 2 4 6 8 100

2 104

4 104

6 104

8 104

1 105

1.2 105

Inte

nsity

Elution time (min)

Br Br

Br

m/z = 578.608

m/z = 500.698

m/z = 328.764

m/z = 516.693

m/z = 498.682

OBrOH

BrBr Br

Page 56: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  51  

Figure A7- Extracted ion mass chromatograms for OH-PBDE 118. All ions were observed as [M-H]-.

OBrOH

BrBr

OH

Br

Br

OH

Br4

OOH

Br4

OH

OOH

Br4

OH

BrBr

Br

0 2 4 6 8 100

5 105

1 106

1.5 106

2 106

2.5 106

3 106

Inte

nsity

Elution time (min)

0 2 4 6 8 100

1 1052 1053 1054 1055 1056 1057 105

Inte

nsity

Elution time (min)

0 2 4 6 8 100

2 105

4 105

6 105

8 105

1 106

Inte

nsity

Elution time (min)

0 2 4 6 8 100

1 104

2 104

3 104

4 104

5 104

Inte

nsity

Elution time (min)

m/z = 578.608

m/z = 500.698

m/z = 328.764

m/z = 516.693

Page 57: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 52  

Figure A8- Plots of appearance vs. time for selected photoproducts of 6’-OH-PBDE 100. Note the Y-axis is labeled “Intensity” for photoproducts that were only identified by LC-MS and concentration is shown when quantification was possible.

0

1 106

2 106

3 106

4 106

5 106

0 1 2 3 4 5 6 7

di-OH-PBDE from 6'-OH-PBDE 100

Inte

nsity

Time (min)

0

2 105

4 105

6 105

8 105

1 106

1.2 106

0 1 2 3 4 5 6 7

di-OH-PBB from 6'-OH-PBDE 100

Inte

nsity

Time (min)

0

0.1

0.2

0.3

0.4

0.5

0.6

0 1 2 3 4 5 6 7

2,4,6 tribromophenol from 6'-OH-PBDE 100

Con

cent

ratio

n [µ

M]

Time (min)

Page 58: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  53  

Figure A9- Proposed mechanism for the formation of hydroxylated dibenzo-p-furan from 6’-OH-PBDE 100.

Br

OO

Br

Br

BrBrBrO

Br

O

BrBr

Br

OO

Br

Br

BrBrBr

OO

Br

Br

BrBrH Br

OO

Br

Br

BrH Br

– Br–OOH

Br

Br

BrBr

6 π e-conrotatory

10 π e-suprafacial

1,3-migration

Page 59: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 54  

Synthesis of compounds 1-3 General All reagents were purchased from commercial suppliers. Solvents were used as received unless otherwise specified. Reactions were performed under ambient conditions unless otherwise stated. Proton nuclear magnetic resonance spectroscopy (1H NMR) and carbon nuclear magnetic resonance spectroscopy (13C NMR) utilized a Varian Inova 500 MHz NMR spectrometer. Fluorine nuclear magnetic resonance spectroscopy (19F NMR) utilized a Varian Unity 300 MHz NMR spectrometer. Electrospray ionization mass spectrometry (ESI-MS) and electrospray ionization high-resolution mass spectrometry (ESI-HRMS) utilized a BIO-TOF mass spectrometer. Melting points were determined manually using a Büchi Schmelzpunktbestimmungs apparat. 2,2’,4,4’-Tetrabromodiphenyliodonium chloride was synthesized according to a literature procedure1.

3-Bromo-2-methoxybenzaldehyde: 3-Bromosalicylaldehyde (0.8854 g, 4.40 mmol), NaOH (0.595 g, 14.9 mmol), and tetrabutylammonium hydroxide (7.16 g, 8.95 mmol) were partitioned between CH2Cl2 (40 mL) and water (40 mL). To this bright yellow solution was added CH3I (1.4 mL, 22 mmol). The solution was allowed to stir at room temperature until colorless, upon which the layers were separated. The aqueous layer was extracted with CH2Cl2 (3 x 50 mL) and the organic layers were combined and dried over Na2SO4. The solvents were removed in vacuo. Flash column chromatography (5:1 CH2Cl2:hexanes; Rf = 0.57) gave the desired product (0.877 g, 93%) as a yellow oil. 1H NMR (500 MHz, CDCl3): δ = 10.31 (s, 1H, CHO), 7.77 (dd, J = 1.5, 7.5 1H, 4-H), 7.75 (dd, J = 1.5, 7.5, 1H, 6-H), 7.10 (dd, J = 7.8, 7.8, 1H, 5-H), 3.95 (s, 3H, CH3) ppm. 13C NMR (500 NMR, CDCl3) δ = 189.02 (CHO), 160.10 (2-C), 139.47 (4-C), 130.94 (1-C), 127.87 (6-C), 125.78 (5-C), 118.19 (3-C), 63.51 (OCH3). ESI-MS: m/z [ion] (rel. int%) 238.8 [M+2] (100), 326.8 [M] (94), 239.8 [M+3] (18), 237.8 [M+1] (12). ESI-HRMS: calculated for [C8H7NaO2Br]+ 236.9527, found 236.9516.

3-Bromo-2-methoxyphenol: 3-Bromo-2-methoxybenzaldehyde (0.768 g, 3.57 mmol) and KH2PO4 (10.0 g, 73.6 mmol) were suspended in CH2Cl2 (40 mL). In another flask was added H2O2 (0.606 g, 5.35 mmol, 30%) and CH2Cl2 (5 mL). The peroxide solution was cooled to 0o C, upon which trifluoroacetic acid anhydride (3.8 mL, 27 mmol) was added dropwise and then allowed to stir for an additional hour at 0o C. The aldehyde mixture was

Page 60: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  55  

cooled to 0o C and the peroxide solution was added to the aldehyde mixture dropwise. The reaction mixture was allowed to stir for 2 h, upon which brine (10 mL) and aqueous NaHSO3 (10 mL, 20%) were added to quench the reaction. The layers were separated and the aqueous layer was extracted with CH2Cl2 (3 x 30 mL). The organic layers were combined and the solvent was removed in vacuo. The residue was dissolved in CH3OH (20 mL) with one drop of concentrated HCl. The reaction mixture was allowed to stir for 12 h and then the solvents were removed in vacuo. Flash column chromatography (4:1 CH2Cl2:hexanes; Rf = 0.24) gave the desired product (0.602 g, 84%) as a yellow oil. 1H NMR (500 NMR, CDCl3): δ = 7.07 (dd, J = 3.0, 8.0, 1H, 4-H), 6.93 (dd, J = 3.0, 8.0, 1H, 6-H), 6.90 (dd, J = 8.0, 8.0, 1H, 5-H), 6.04 (m, 1H, OH), 3.90 (s, 3H, CH3) ppm. 13C NMR (500 NMR, CDCl3) δ = 150.10 (1-C), 144.52 (2-C), 126.02 (5-C), 124.70 (4-C), 116.01 (3-C), 115.24 (6-C), 61.15 (CH3) ppm. ESI-MS: m/z [ion] (rel. int%) 185.9 [M-1] (100), 187.9 [M+1] (98), 188.9 [M+3] (10), 186.9 [M] (6). ESI-HRMS: calculated for [C7H6BrO2]- 200.9551, found 200.9554.

3,4,6-Tribromo-2-methoxyphenol: 3-Bromo-2-methoxyphenol (0.435 g, 2.14 mmol) and CaCO3 (0.644 g, 4.71 mmol) were suspended in CH2Cl2 (50 mL) and CH3OH (20 mL). Benzyltrimethylammonium tribromide (1.84 g, 6.43 mmol) was added in small portions over 2 h and then the reaction mixture was allowed to stir for an additional 2 h. The reaction mixture was filtered and the filtrate was added to aqueous NaHSO3 (40 mL, 5%). The layers were separated and the aqueous layer was extracted with CH2Cl2 (3x 50 mL). The organic fractions were combined and dried over Na2SO4. The solvents were removed in vacuo. Flash column chromatography (4:1 CH2Cl2:hexanes; Rf = 0.40) gave the desired product (0.702 g, 91%) as light yellow solid. 1H NMR (500 NMR, CDCl3) δ = 7.58 (s, 1H, 5-H), 5.97 (s, 1H, OH), 3.92 (s, 3H, CH3) ppm. 13C NMR (500 NMR, CDCl3) δ = 146.75 (2-C), 146.09 (1-C), 131.34 (5-C), 119.23 (4-C), 115.46 (3-C), 108.99 (6-C), 61.16 (OCH3). ESI-MS: m/z [ion] (rel. int%) 358.8 [M+1]- (100), 360.8 [M+3]- (96), 356.8 [M-1] - (24), 362.8 [M+5] - (24). ESI-HRMS: calculated for [C7H4Br3O2]- 358.7741, found 358.7741.

1,2,5-tribromo-4-(2,4-dibromophenoxy)-3-methoxybenzene: 3,4,6-Tribromo-2-methoxyphenol (0.334 g, 0.93 mmol), K2CO3 (0.339 g, 2.45 mmol), and 18-crown-6 ether (0.034 g, 0.13 mmol) were suspended in N,N-dimethylacetamide (25 mL). 2,2’,4,4’-Tetrabromo-diphenyliodonium chloride (0.800 g, 1.27 mmol) was added to the reaction mixture and then heated to 80 oC where it

Page 61: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 56  

was stirred for 1 h. The mixture was cooled to room temperature and diluted with CH2Cl2 (30 mL) and water (30 mL). The layers were separated and the aqueous layer was extracted with CH2Cl2 (3 x 50 mL). The organic fractions were combined, washed with aqueous NaHSO3 (50 mL, 5%), aqueous NaOH (2 x 50 mL, 1.0 M), water (3 x 50 mL), and dried over Na2SO4. The solvent was removed in vacuo. Flash column chromatography (4:1 hexanes:CH2Cl2, Rf = 0.56) gave the desired product (0.355 g, 64%) as a colorless resin. 1H NMR (500 MHz, CDCl3): δ = 7.77 (d, J = 2.0, 1H, 3’-H), 7.75 (s, 1H, 4-H), 7.26 (dd, J = 2.5, 9.0, 1H, 5’-H), 6.33 (d, J = 9.0, 1H, 6’-H), 3.86 (s, 3H, CH3) ppm. 13C NMR (500 NMR, CDCl3) δ = 152.53 (1-C), 152.34 (1’-C), 144.96 (2-C), 136.01 (3’-C), 132.01 (5’-C), 131.38 (4-C), 122.77 (5-C), 121.40 (3-C), 117.36 (4’-C), 115.59 (6’-C), 115.42 (6-C), 112.05 (2’-C), 61.82 (OCH3). ESI-MS: m/z [ion] (rel. int%) 618.5 [M+4] (100), 616.5 [M+2] (82), 614.5 [M] (57), 620.5 [M+6] (35). HRMS: calculated for [C13H7Br5NaO2]+ 618.6199, found 618.6184

2,3,5-tribromo-6-(2,4-dibromophenoxy)phenol : 1,2,5-tribromo-4-(2,4-dibromophenoxy)-3-methoxybenzene (0.147 g, 0.24 mmol) was dissolved in a round bottom flask equipped with a reflux condenser with dry CH2Cl2 (15 mL, from CaH2). BBr3 (240 mL, 2.4 mmol) was added and the solution was heated to reflux for 48 h, upon which the solution was cooled to 0 oC and water (15 mL) was carefully added. The solution was extracted with CH2Cl2 (3 x 25 mL) and dried over Na2SO4. The solvents were removed in vacuo. Flash column chromatography (1:1 CH2Cl2:hexanes, Rf = 0.26) gave the desired product (0.132 g, 92%) as a colorless resin. 1H NMR (500 MHz, CDCl3): δ = 7.78 (d, J = 2.5, 1H, 3’-H), 7.55 (s, 1H, 4-H), 7.28 (dd, J = 2.5, 8.5, 1H, 5’-H), 6.42 (d, J = 8.5, 1H, 6’-H), 6.05 (s, 1H, OH) ppm. 13C NMR (500 NMR, CDCl3) δ = 152.18 (1’-C), 148.21 (1-C), 138.720 (2-C), 136.15 (3’-C), 131.39 (5’-C), 128.05 (4-C), 122.24 (5-C), 116.78 (3-C), 115.84 (6’-C), 113.50 (4’-C), 113.21 (6-C), 112.67 (2’-C). ESI-MS: m/z [ion] (rel. int%) 578.6 [M+3] (100), 580.6 [M+5] (93), 576.6 [M+1] (40), 582.6 [M+7] (36), 574.6 [M] (5), 584 [M+9] (4). ESI-HRMS: calculated for [C12H4Br5O2]- 578.6087, found 578.6074.

4,5-Dibromo-guaiacol: Guaiacol (1.02 g, 8.18 mmol) was added to a round bottom flask equipped with a magnetic stir bar and an addition funnel. After purging the reaction vessel with dry nitrogen gas, bromine (0.84 mL, 16 mmol) dissolved in CH2Cl2 (4 mL) was added to the guaiacol over the course of 2.5 h. After complete addition of the bromine solution, the solvent was removed under reduced pressure to yield the crude product as a white

Page 62: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  57  

solid. Recrystallization from EtOH/water (2:3) resulted in the desired product (1.945 g, 84%) as long white needles. 1H NMR (500 NMR, CDCl3): δ = 7.17 (s, 1H, 6-H), 7.05 (s, 1H, 3-H), 5.61 (s, 1H, OH), 3.88 (s, 3H, OCH3) ppm. 13C NMR (500 NMR, CDCl3) δ = 146.45 (2-C), 145.54 (1-C), 119.14 (6-C), 115. 47 (5-C), 115.36 (3-C), 113.80 (4-C) ppm. ESI-MS: m/z [ion] (rel. int%) 280.9 [M+1] (100), 282.9 [M+3] (45), 278.9 [M-1] (44), 281.9 [M+2] (4), 279.9 [M] (3), 283.9 [M+4] (2). ESI-HRMS: calculated for [C7H5Br2O2]- 280.8636, found 280.8637. Melting point: 92-93 oC, lit. 94-95 oC (2).

2,5-Dibromo-4-nitro-fluorobenzene: 2,5-Dibromo-fluorobenzene (5.03 g, 19.8 mmol) was placed in a round bottom flask equipped with a magnetic stir bar. The starting material was dissolved in CH2Cl2 (20 mL), trifluoroacetic anhydride (10 mL), and trifluoroacetic acid (20 mL). The solution was cooled to 0 oC and to this solution was added ammonium nitrate (1.98 g, 24.7 mmol). The clear orange solution was allowed to warm to room temperature overnight, upon which the solvent was removed via rotary evaporation. Flash column chromatography (1:1 CH2Cl2:hexanes; Rf = 0.60) provided the desired product (5.04 g, 85%) as a light yellow solid. 1H NMR (500 MHz, CDCl3): δ = 8.528 (d, J = 6.5, 1H, 3-H), 8.11 (d, J = 8.5, 1H, 6-H) ppm. 13C NMR (500 MHz, CDCl3) δ = 160.22 (d, J = 255.7, 1-C), 147.12 (s, 4-C), 130.77 (s, 3-C), 123.14 (d, J = 28.2, 6-C), 114.53 (d, J = 10.1, 5-C), 108.95 (d, J = 23.1, 2-C) ppm. 19F NMR (300 MHz, CDCl3) δ = -99.79 (1-F) ppm.

1,4-dibromo-2-(4,5-dibromo-2-methoxyphenoxy)-5-nitrobenzene: 2,5-Dibromo-4-nitro-fluorobenzene (1.08 g, 3.6 mmol) and K2CO3 (1.49 g, 10.8 mmol) were placed a 2-necked round bottom flask equipped with a magnetic stir bar, reflux condenser, and an addition funnel with pressure equalizing sidearm containing 4,5-dibromo-guaicol (1.02g, 3.6 mmol). The reaction set-up was placed under vacuum for 1 h, after which it was purged with nitrogen gas. Dry acetone was added to the reaction flask (10 mL) and to the addition funnel (6 mL). The reaction flask was heated to reflux and then the guaicol solution was slowly added over the course of 1 h. The reaction was allowed to stir for an additional 2 h upon completion of the addition and then cooled to room temperature. The solvent was removed via rotary evaporator and the resulting residue was partitioned in CH2Cl2:water (1:1). The layers were separated and the aqueous layer was extracted with CH2Cl2 (3 x 50 mL). The combined organic layers were washed with 10% NaOH, brine, and then dried over anhydrous Na2SO4. Removal of

Page 63: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 58  

the solvent via rotary evaporator left a viscous red oil. Flash column chromatography (1:1 CH2Cl2:hexanes; Rf = 0.52) and subsequent recrystallization from EtOH/water yielded the desired product (1.26 g, 62%) as fine white needles. 1H NMR (500 MHz, CDCl3): δ = 8.26 (s, 1H, 6-H), 7.40 (s, 1H, 3-H), 7.30 (s, 1H, 6’-H), 6.84 (s, 1H, 3’-H), 3.80 (s, 3H, OCH3) ppm. 13C NMR (500 MHz, CDCl3) δ = 157.45 (2-C), 150.80 (2’-C), 144.06 (5-C), 141.59 (1’-C), 131.18 (6-C), 126.84 (3-C), 122.62 (4’-C), 120.60 (6’-C), 118.15 (3’-C), 115.40 (5’-C), 115.07 (1-C), 110.79 (4-C), 56.58 (OCH3) ppm. ESI-MS: m/z [ion] (rel. int%) 585.4 [M+6] (100), 583.4 [M+4] (86), 581.4 [M+2] (50), 586.4 [M+8] (27). ESI-HRMS: calculated for [C13H7Br4NNaO4]+ 585.6945, found 585.6945.

2,5-dibromo-4-(4,5-dibromo-2-methoxyphenoxy)-benzenaminium chloride: 1,4-dibromo-2-(4,5-dibromo-2-methoxyphenoxy)-5-nitrobenzene (0.168 g, 0.30 mmol), CH3OH (50 mL), and glacial acetic acid (5 mL) were placed in a round bottom flask equipped with a magnetic stir bar and reflux condenser. To this solution was added iron powder (0.515 g, 9.22 mmol, ~325 mesh) and the solution was heated to reflux. After 5 h at reflux, the solvent was removed via rotary evaporator. The residue was suspended in benzene and the iron was removed via vacuum filtration. The filtrate was washed with boiling benzene (3 x 50 mL) and then the combined organic fractions were cooled to 0 oC. Dry HCl (HCl on anhydrous calcium chloride) was bubbled through the cooled benzene solution to precipitate out the desired product (0.118 g, 69%) as a pale solid. 1H NMR (500 MHz, d6-DMSO): d = 7.46 (s, 1H, 6-H), 7.18 (s, 1H, 3-H), 7.12 (s, 1H, 6’-H), 6.83 (s, 1H, 3’-H), 3.92 (br, 3H, NH3

+), 3.83 (s, 3H, OCH3) ppm. ESI-MS: m/z [ion] (rel. int%) 531.7 [M+5] (100), 529 [M+2] (61), 533.7 [M+7] (51). ESI-HRMS: calculated for [C13H10Br4NO2]+ 531.7404, found 531.7395. The corresponding free amine was isolated following alkaline workup. 1H NMR (500 MHz, CDCl3): δ = 7.18 (s, 1H, 6’-H), 7.07 (s, 1H, 3’-H), 7.02 (s, 1H, 3-H), 6.86 (s, 1H, 6-H), 4.09 (s, 2H, NH2), 3.88 (s, 3H, OCH3) ppm. 13C NMR (500 MHz, CDCl3) δ = 149.52 (2’-C), 146.48 (1’-C), 144.10 (4-C), 142.36 (1-C), 124.79 (3-C), 121.22 (6’-C), 119.25 (3’-C), 117.97 (4’-C), 117.37 (6-C), 114.81 (5’-C), 114.69 (2-C), 107.89 (5-C), 56.61 (OCH3) ppm. ESI-MS: m/z [ion] (rel. int%) 553.7 [M+4] (100), 555.7 [M+6] (71), 551.7 [M+2] (60), 549.7 [M] (19), 557.7 [M+7] (18), 554.7 [M+5] (13), 552.7 [M+3] (12), 556.7 [M+7] (12), 550.7 [M+1] (2). ESI-HRMS: calculated for [C13H9Br4NNaO2]+ 553.7224, found 553.7224.

Page 64: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  59  

1,2,4-tribromo-5-(4,5-dibromo-2 methoxyphenoxy)benzene: Copper(II) bromide (0.046 g, 0.20 mmol) was placed in a two-necked round bottom equipped with a magnetic stir bar, reflux condenser, and an addition funnel. The reaction apparatus was purged with nitrogen, and tert-butyl nitrite (30 mL, 0.25 mmol) and dry acetonitrile (10 mL) were added to the reaction flask. 2,5-Dibromo-4-(4,5-dibromo-2-methoxyphenoxy)aniline (0.088 g, 0.17 mmol) was dissolved in CH2Cl2 (5 mL) and charged into the addition funnel. The reaction mixture was heated to reflux and the amine solution was added dropwise over 5 min. The reaction solution was allowed to stir at reflux for 3 h, after which it was cooled to room temperature and poured into 20% aqueous HCl (75 mL). The resulting white precipitate was extracted with Et2O (2 x 100 mL). The combined organic fractions were washed with 20% aqueous HCl (50 mL) and then dried over anhydrous Na2SO4. Excess solvent was removed via rotary evaporator and flash column chromatography (4:1 hexanes: CH2Cl2; Rf = 0.38) yielded the desired product (0.081 g, 82%) as a white solid. 1H NMR (500 MHz, CDCl3): δ = 7.85 (s, 1H, 5-H), 7.25 (s, 1H, 2-H), 7.21 (s, 1H, 6-H), 6.92 (s, 1H, 3-H), 3.82 (s, 3H, OCH3) ppm. 13C NMR (500 MHz, CDCl3): δ = 153.54 (1-C), 150.67 (2’-C), 143. 29 (1’-C), 137.11 (5-C), 125.32 (2-C), 124.01 (3-C), 121.67 (6’-C), 120.96 (4’-C), 118.90 (4-C), 117.90 (3’-C), 115.11 (5’-C), 112.39 (6-C) ppm.

4,5-dibromo-2-(2,4,5-tribromophenoxy)phenol: 1,2,4-tribromo-5-(4,5-dibromo-2-methoxyphenoxy)benzene (0.041 g, 0.07 mmol) was dissolved in a round bottom flask equipped with a reflux condenser with dry CH2Cl2 (10 mL, from CaH2). BBr3 (66 mL, 0.70 mmol) was added and the solution was heated to reflux for 72 h, after which time the solution was cooled to 0 oC and water (10 mL) was carefully added. The solution was extracted with CH2Cl2 (3 x 25 mL) and dried over Na2SO4. The solvents were removed via rotary evaporator. Flash column chromatography (1:1 CH2Cl2:hexanes, Rf = 0.21) gave the desired product (0.034 g, 85%) as a white resin. 1H NMR (500 MHz, CDCl3): δ = 7.90 (s, 1H, 5’-H), 7.35 (s, 1H, 2’-H), 7.23 (s, 1H, 3-H), 6.97 (s, 1H, 6-H), 5.74 (s, 1H, OH) ppm. 13C NMR (500 NMR, CDCl3) δ = 151.84 (1’-C), 146.49 (1-C), 142.74 (2-C), 137.58 (3’-C), 124.63 (5’-C), 124.46 (6’-C), 123.38 (5-C), 121.84 (3-C), 121.52 (6-C), 120.25 (4’-C), 114.63 (4-C), 113.90 (2’-C) ppm. ESI-MS: m/z [ion] (rel. int%) 578.7 [M+3]

Page 65: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 60  

(100), 580.7 [M+5] (91), 576.7 [M+1] (39), 582.7 [M+7] (29), 574.7 [M-1] (6), 584 [M+9] (4). ESI-HRMS: calculated for [C12H4Br5O2]- 578.6087, found 578.6087.

2,4,6-Tribromo-iodobenzene: To a slurry of 2,4,6-tribromoaniline (15.0 g, 45.5 mmol) suspended in concentrated HCl (30 mL) cooled to 0 oC was slowly added an aqueous solution of NaNO2 (3.28 g, 47.5 mmol, 15 mL). The resulting yellow solution was filtered through glass wool and carefully added to an aqueous solution of KI (75.6 g, 455 mmol, 115 mL) at room temperature. The mixture was allowed to stir for 1 h upon completion of the addition, after which CH2Cl2 (150 mL) and aqueous Na2SO3 (1.0 M, 10 mL) was added. The layers were separated and the aqueous layer was extracted with CH2Cl2 (3 x 100 mL). The combined organic layers were washed with 10% HCl, brine, and dried over Na2SO4. Removal of solvent via rotary evaporator left a crude red solid. Recrystallization from CH2Cl2:hexanes (3:1) yielded the desired product (6.76 g, 34%) as pale white crystals. 1H NMR (500 MHz, CDCl3): δ = 7.70 (s, 2H, 3-H, 5-H). 13C NMR (500 MHz, CDCl3) δ = 133.57 (2C, 3-C, 5-C), 131.85 (2C, 2-C, 6-C), 122.98 (1C, 4-C), 108.18 (1C, 1-C) ppm. Melting point: 104-105oC, lit. 104-105 oC3.

(4-methoxyphenyl)(2,4,6-tribromophenyl)iodonium bromide: H2O2 (222 mL, 1.96 mmol, 30%) was placed in a pear-shaped flask equipped with a stir bar and cooled to 0 oC. Trifluoroacetic anhydride (1.85 mL, 13.3 mmol) was added over 5 min followed by 2,4,6-tribromo-iodobenzene (1.10 g, 2.50 mmol) dissolved in CH2Cl2 (5 mL) added over 10 min. The solution was allowed to warm up to room temperature overnight after which the solvents were removed in vacuo. The resulting pale solid was dissolved in 10 mL of CH2Cl2:acetic anhydride (1:1) and glacial acetic acid (192 mL) was added. The solution was cooled to -20 oC and anisole (550 mL, 5.04 mmol) dissolved in CH2Cl2 (5 mL) was added dropwise. After the addition was complete, the solution was allowed to stir at -20 oC for 1 h and then warmed to room temperature and stirred for another 1 h. The resulting light blue solution was placed under vacuum to remove the solvent and the residue was dissolved in a minimal amount of CH3OH (10 mL). A saturated solution of aqueous NaBr (5 mL) was added to the methanolic solution and the resulting precipitate was filtered off and dried in a vacuum oven to yield the desired product (1.51 g, 96%) as a light yellow solid (purity >90% via 1H NMR). 1H NMR (500 MHz, d6-DMSO): δ = 8.12 (s, 2H, 3-H, 5-H), 7.99 (d, J = 8.5, 2H, 2’-H, 6’-H), 7.02 (d, J = 8.7, 2H, 3’-H, 5’-H) 3.75 (s, 3H, OCH3) ppm. 13C NMR (500 MHz, d6-DMSO): δ = 161.78 (C’OCH3),

Page 66: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  61  

134.61 (2C, 2’-C, 6’-C), 133.73 (1-C), 129.78 (4-C), 129.27 (2C, 3-C, 5-C), 117.50 (2C, 3’-C, 5’-C), 114.13 (1’-C), 110.87 (2C, 2-C, 6-C), 55.90 (OCH3) ppm.

3,5-dibromo-2-(2,4,6-tribromophenoxy)benzaldehyde: 3,5-Dibromo-salicylaldehyde (0.341 g, 1.22 mmol) was placed in a round bottom flask equipped with a magnetic stir bar and a reflux condenser. The aldehyde was dissolved in 20 mL of dioxane:water (1:1) and then to the solution was added NaOH (0.102 g, 2.54 mmol). To the resulting bright yellow solution was added (4-methoxyphenyl)(2,4,6-tribromophenyl)iodonium bromide (1.14 g, 1.81 mmol) and the reaction mixture was heated to 80 oC. After 2 h at 80 oC, ice was added to the solution and the resulting mixture was extracted with CH2Cl2 (3 x 50 mL). The organic fractions were washed with 1 M aqueous NaOH (2 x 50 mL), water (2 x 50 mL), and then dried over Na2SO4. The solvent was removed via rotary evaporator and flash chromatography (3:2 hexanes: CH2Cl2; Rf = 0.64) yielded the desired product (0.302 g, 42%) as a white solid. 1H NMR (500 MHz, CDCl3): δ = 10.37 (s, 1H, CHO), 7.98 (s, 1H, 4-H), 7.87 (s, 1H, 2-H), 7.69 (s, 2H, 3’-H, 5’-H) ppm. 13C NMR (500 MHz, CDCl3): δ = 186.96 (CHO), 153.93 (2-C), 149.18 (1’-C), 141.83 (4-C), 135.63 (2C, 3’-C, 5’-C), 130.94 (6-C), 130.19 (1-C), 118.15 (4’-C), 117.93 (2C, 2’-C, 6’-C), 116.37 (3-C), 113.48 (5-C) ppm. ESI-MS: m/z [ion] (rel. int%) 614.5 [M+4] (100), 616.5 [M+6] (95), 612.5 [M+2] (56), 618.5 [M+8] (47), 610.5 [M] (13), 620.5 [M+10] (8). ESI-HRMS: calculated for [C13H5Br5NaO2]+ 614.6063, found 614.6067.

3,5-dibromo-2-(2,4,6-tribromophenoxy)phenol: 3,5-dibromo-2-(2,4,6-tribromophenoxy)benzaldehyde (0.131 g, 0.22 mmol) and KH2PO4 (0.240 g, 1.76 mmol) were placed in a round bottom flask equipped with a magnetic stir bar and suspended in CH2Cl2 (10 mL). In another flask was added H2O2 (0.050 g, 0.44 mmol, 30%) and CH2Cl2 (1 mL). The peroxide solution was cooled to 0 oC, to which trifluoroacetic acid anhydride (230 mL, 1.65 mmol) was added dropwise and then allowed to stir for 1 h at 0 oC. The aldehyde mixture was cooled to 0 oC and then the peroxide solution was added dropwise. The reaction mixture was allowed to stir for 5 h, upon which brine (20 mL) and aqueous NaHSO3 (20 mL, 20%) were added to quench the reaction. The layers were separated and the aqueous layer was extracted with CH2Cl2 (3 x 30 mL). The organic layers were combined and the solvent was removed in vacuo. The residue was dissolved in CH3OH (20 mL) with one drop of concentrated HCl. The reaction mixture was allowed to stir for 18 h and

Page 67: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 62  

then the solvents were removed via rotary evaporator. Flash column chromatography (4:1 CH2Cl2:hexanes; Rf = 0.42) gave the desired product (0.112 g, 87%) as a white solid. 1H NMR (500 NMR, CDCl3): δ = 7.68 (s, 2H, 3’-H, 5’-H), 7.19 (d, J = 2.5, 1H, 4-H), 7.12 (d, J = 2.0, 1H, 6-H), 5.84 (s, 1H, OH) ppm. 13C NMR (500 NMR, CDCl3) δ = 149.19 (1-C), 135.63 (1’-C), 135.40 (2C, 3’-C, 5’-C), 127.58 (4-C), 118.88 (6-C), 117.92 (2-C), 117.51 (3-C), 116.65 (4’-C), 113.10 (2C, 2’-C, 6’-C), 111.78 (5-C) ppm. ESI-MS: m/z [ion] (rel. int%) 578.7 [M+3] (100), 580.7 [M+5] (85), 576.7 [M+1] (44), 582.7 [M+7] (34), 574.7 [M-1] (7), 584 [M+9] (6). ESI-HRMS: calculated for [C12H4Br5O2]- 578.6087, found 578.6087.

Page 68: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  63  

OHBr

O OMeBr

O OMeBrHO

OMeBrHO

BrI+

+Br

Br

Br

Cl- OO

Br

Br

Br

a

Br

Br

e

OHO

Br

Br

Br

Br

Br

b c

d

Figure S10- Synthesis of OH-PBDE-99. Reaction conditions: a) (C4H9)4NOH, NaOH, MeI, CH2Cl2, H2O, 93%; b) i) H2O2, (CF3CO)2O, KH2PO4, CH2Cl2, ii) MeOH, HCl, 84%; c) BTMA-Br3, CaCO3, MeOH, CH2Cl2, 91%; d) K2CO3, 18-crown-6, DMAC, 64%; e) BBr3, CH2Cl2, 92%.

Br Br

BrNH2

BrI HO

+a

O

Br BrBr

b c

Figure S11- Synthesis of OH-PBDE-100. Reaction conditions: a) i) NaNO2, HCl, 0oC, ii) KI, H2O 34%; b) i) H2O2, TFA, TFAA, CH2Cl2, ii) Ac2O, CH3CO2H, anisole, CH2Cl2, iii) NaBr, MeOH, H2O, 96%; c) NaOH, dioxane, H2O, 42%; d) i) H2O2, (CF3CO)2O, KH2PO4, CH2Cl2, ii) MeOH, HCl, 87%.

Br Br Br Br

BrI+

Br O

Br-

Br Br

O

Br

Br

Br O

O

Br

OH

BrBr Br

Brd

F a

b

cBr

MeO

F

Br

O2N Br

MeO Br

O

O2N

OMe

Br

Br

O

H2N

OMe

Br

BrO

Br

OMe

Br

BrO

Br

OH

Br

Br

Br Br

HO HOBrBr

d

e f

BrBr BrBr BrBr

Figure S12- Synthesis of OH-PBDE-118. Reaction conditions: a) NH4NO3, TFAA, TFA, CH2Cl2, 85%; b) Br2, CH2Cl2, 84%; c) K2CO3, acetone, 62%; d) Fe, MeOH, HOAc, 69%; e) CuBr, tBuNO, CH3CN, 82%; f) BBr3, CH2Cl2, 85%.

Page 69: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 64  

Synthesis of dibenzofuran derivatives General All reagents were purchased from commercial suppliers. Solvents were used as received unless otherwise specified. Reactions were performed under nitrogen. Proton nuclear magnetic resonance spectroscopy (1H NMR) and carbon nuclear magnetic resonance spectroscopy (13C NMR) were conducted with a Bruker Avance III 400 MHz NMR spectrometer. Exact mass measurements were acquired by ESI-HRMS using a Thermo Exactive Orbitrap instrument. Dibenzofuran-4-ol was prepared following a previously published method4. Carbon numbers were assigned following IUPAC nomenclature for fused rings5.

Dibenzofuran-4-ol: Dibenzofuran (2.02 g, 12.0 mmol), TMEDA (2.3 g, 19.8 mmol), and dry Et2O (50 mL) were placed in a 2-neck round bottom flask equipped with a reflux condenser and stir bar. With stirring, n-BuLi (5 ml, 2.5 M in hexane) was added and the mixture was heated to reflux for 1 h. The mixture was cooled to 0 °C and tributyl borate (3.8 mL, 14.4 mmol) was added until the yellow precipitate dissolved. The mixture was kept at 0 °C for 45 min, and then was allowed to warm to room temperature over 1 h. The mixture was again cooled to 0 °C and 30% H2O2 (5 ml) was added dropwise while stirring. The mixture was heated to reflux for 1.5 h, cooled to 0 °C and acidified with 5 M HCl (10 mL). The organic phase was washed with cold 10% ammonium iron(II) sulphate (2 x 50 mL) and extracted with 2 M NaOH (3 x 50 mL). The combined aqueous extract was acidified and extracted with Et2O (3 x 100 mL) and dried with Na2SO4. The solvent was evaporated in vacuo and the crude mixture was filtered through silica gel with 15% Et2O:petroleum ether giving the desired product (1.08 g, 48%) as off-white crystals. 1H NMR (400 MHz, CDCl3) δ= 7.94 (dd, J = 0.7, 7.7 1H, 9-H), 7.59 (d, 1H, 6-H), 7.52 (dd, 1H, 1-H), 7.47 (td, 1H, 7-H), 7.36 (td, 1H, 8-H), 7.22 (t, 1H, 2-H), 7.02 (d, 1H. 3-H), 5.33 (s, 1H, OH). 13C NMR (400MHz, CDCl3) δ= 156.21 (4-C), 144.18 (4a or 5a-C), 141.27 (4a or 5a-C), 127.43 (7-C), 125.88 (9a or 9b-C), 124.75 (9a or 9b-C), 123.83 (2-C), 123.14 (8-C), 121.18 (9-C), 113.69 (3-C), 112.96 (1-C), 111.94 (6-C) ESI-HRMS: calculated for [C12H7O2]- 183.0446, found 183.0446.

1,2,3,8-tetrabromo-dibenzofuran-4-ol: Dibenzofuran-4-ol (.52 g, 2.83 mmol) was placed in a 2-neck round bottom flask equipped with an addition funnel and stir bar and dissolved in a minimum amount of dry CH2Cl2 (10mL). To the addition funnel was added a mixture of Br2 (2.79 g, 17.4 mmol) and dry CH2Cl2 (4 mL). The Br2 solution was added slowly over a period of 15 min at room temperature. After the addition was complete, the mixture was stirred at

OOH

OOH

Br

BrBr

Br

Page 70: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  65  

room temperature for 30 h. The remaining Br2 was removed by washing the mixture with a solution of 10% Na2SO3 until the red color disappeared. The mixture was then extracted with CH2Cl2 (3 x 100 mL), dried with Na2SO4, and the solvent was removed in vacuo. The crude product only contained a small fraction of the desired product, which was collected as the last fractions on a flash column (80:20 hexane: ethyl acetate). 1H NMR (400 MHz, CDCl3) δ = 8.63 (dd, J = 2.0, 1H, 9-H), 6.67 (dd, J = 2.0, 8.8, 1H, 7-H), 7.51 (d, J =8.8, 1H, 6-H), 6.02 (br s, 1H, OH). ESI-HRMS: calculated for [C12H3O2Br4]- 498.6826, found 498.6827. References:

1) Steen, Peter O.; Grandbois, Matthew; McNeill, Kristopher; Arnold, William A.; Photochemical Formation of Halogenated Dioxins from Hydroxylated Polybrominated Diphenyl Ethers (OH-PBDEs) and Chlorinated Derivatives (OH-PBCDEs). Environmental Science and Technology 2009, 43, 4405-4411.

2) Raiford, L. C.; Silker, Ralph E.; Nitration of Acyl Derivatives of 4,5-Dibromo- and 4,5,6-tribromoguaiacol. Journal of Organic Chemistry 1937, 2, 346-355.

3) Bolton, Roger; Sandall, John P. B.; Nucleophilic displacement in polyhalogenoaromatic compounds. Part 3. Kinetics of protiodeiodination of iodoarenes. Journal of Chemical Society, Perkins Transactions 2 1977, 278-280.  

4) Oliveira, A. M. A. G.; Raposo, M. M. M.; Oliveira-Campos, A. M. F.; Griffiths, J.; Machado, A. E. H. Synthesis of Psoralen Analogues Based on Dibenzofuran. Helvetica Chimica Acta 2003, 86, 2900-2907.

5) Fused Ring and Bridged Fused Ring Nomenclature (IUPAC Recommendations 1998) www.chem.qmul.ac.uk/iupac/fusedring/FR53.html#551

Page 71: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 66  

Page 72: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  67  

Chapter 3

Disparate controlling factors in the rates of

oxidation of anilines and phenols by triplet

methylene blue in aqueous solution Paul R. Erickson, Nicolas Walpen, Jennifer Guerard, Soren E. Eustis, Samuel

J. Arey and Kristopher McNeill

In preperation for Journal of Physical Chemistry: A

Page 73: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 68  

Abstract

In aqueous solution, Methylene Blue (MB) in its first excited triplet state can

act as a one-electron oxidant for molecules with a sufficiently low oxidation potential,

such as anilines. Additionally, phenols, which have much higher oxidation potentials,

may also react with Methylene Blue via another mechanism involving a proton-

coupled electron transfer. Here we present a series of experiments showing that the

rate constant of a one-electron transfer from anilines to triplet state Methylene Blue

follows a Sandros-Boltzmann-like dependence on the free energy of the reaction. The

observed rate constants are also well modelled when using aniline oxidation potentials

derived computationally. For phenols, the proton-coupled electron transfer rate

constants were found to depend on O-H bond dissociation free energy with an

additional sensitivity to the phenol pKa. Rate constants could again be modelled using

computed bond dissociation free energies. Furthermore, overall reaction rate constants

for compounds containing both phenolic and amino functional groups could be

predicted by summing the rate constants for the independent mechanisms. These

results show that in situations where rate constant prediction is needed, such as in

aqueous environmental pollutant lifetime prediction, the ability to empirically

determine reaction rate constants may prove valuable.

Page 74: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  69  

Introduction

Anilines and phenols are structurally similar compound classes that play

important roles throughout chemistry and biochemistry. Having electron-rich π

systems, anilines and phenols are prone to oxidation. Our interest in the oxidation of

these compound classes derives from their prevalence in both natural and

anthropogenic molecules that are frequently found in the environment, as well as the

fact that photooxidation is a major environmental degradation process for them1-4.

Despite their structural similarity, they often undergo oxidation by different

mechanisms. Anilines have lower oxidation potentials than phenols, which allows

them to be more easily oxidized by one-electron transfer reactions. Phenols are also

prone to oxidation, but are preferentially oxidized directly to phenoxyl radicals by

hydrogen atom transfer, or related mechanisms.

Given the wide range of possible substituted anilines and phenols, it would be

desirable to be able to predict rate constants for their possible degradation reactions in

the environment with (photo)oxidants. In order to accurately do this, we must first

understand the mechanisms for each class of compound, and also have reliable

parameters to aid in these predictions, such as oxidation potentials in electron transfer

reactions, and bond disassociation enthalpies in the case of hydrogen atom transfer. In

this work, we describe our efforts to determine and accurately model the reaction rate

constants for one-electron transfers from anilines and proton-coupled electron

transfers from phenols to Methylene Blue (MB).

N O HH H

aniline phenol

Page 75: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 70  

Materials and Methods

Chemicals

Methylene Blue (>98.5% spectroscopic grade) was purchased from TCI. All anilines

and phenols used were purchased from Sigma Aldrich or TCI and were of the highest

purity commercially available. D2O (99.9% D) was purchased from Armar Chemicals

(Döttingen, Switzerland). Aqueous solutions were made with ultrapure water (18

MΩ·cm-1) obtained from a Barnstead Nanopure Diamond system. All gasses used

were high purity grade (99.999%).

Transient absorption experiments

The system has been previously described in detail5. Briefly, the setup used was an

EOS transient absorption spectroscopy system (Ultrafast Systems, Sarasota, USA).

Excitation pulses were generated by converting the 795 nm output of a Solstice

regeneratively amplified Ti:sapphire laser (Spectra-Physics, Darmstat, Germany) to

the proper wavelength using a TOPAS-C (Light Conversion, Lithuania) optical

parametric amplifier. The excitation wavelength used was 665 nm. The excitation

pulse energy was 4 µJ or lower. In all experiments, methylene blue solutions (10 µM)

were prepared immediately before analysis and placed in a continuously purged (air

or nitrogen) 10 mm quartz cuvette. All quenching experiments with aniline

derivatives were performed in pH 8.5 borate buffered solutions, and the quencher

concentration ranged from 0.1 to 1.6 mM. Experiments with phenolic quenchers were

performed in water that was adjusted to pH 7 by addition of HCl. Phenolic quencher

concentrations ranged from 0.1 to 2.4 mM. Transient decay lifetimes were calculated

Page 76: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  71  

from fits obtained from Surface Explorer (Ultrafast Systems, Sarasota, USA) and

OrginPro (Version 8.5, Northhampton, USA)

Computational methods

Aqueous oxidation potentials (anilines) and bond dissociation free energies (phenols)

were computed for all compounds within the test set. The loss of an electron in

aqueous phase is defined as:

Aaq → e-g + A•+

(1)

From a computational standpoint it is convenient to use a thermodynamic cycle

(Figure 1A) to separate the free energy of reaction, DG0ox, into the sum of the gas

phase adiabatic ionization energy (IEgas) and the difference in free energy of solvation

between the oxidized and reduced species, DDGsolv:

(2)

(3)

where DG0solv,A is the free energy of solvation of the reduced closed shell neutral

species, and DG0solv,A•+ is the free energy of solvation of the oxidized open shell

cation species. The subscript 0 denotes corresponding to a standard state of 1 atm in

the gas phase and 1 mol/L in solution. The resulting free energy of oxidation, ,

can then be used to determine the oxidation potential:

(4)

where n is the number of electrons, F is the Faraday constant (96,485.3365 C/mol),6

and SHE is the potential of the standard hydrogen electrode, 4.28 V.7 Calculated

DG0ox values were obtained using eqs 2 and 3, based on computed values of IEgas,

!

!

"Gox0 = IEgas(298K )+""Gsolv

!

!

""Gsolv = "G0solv,A•+ # "G0

solv ,A

ΔGox0

!

!

Eox = ""#Gox

0

nF+ SHE

$

% &

'

( )

Page 77: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 72  

DG*solv,A, and DG*

solv,A•+.

Bond dissociation free energies (BDFEs) were modeled as:

(5)

BDFEs were computed according to the thermodynamic cycle shown in Figure A2 B.

Each BDFE was modeled as the sum of the gas phase free energy of reaction

(DGgas,298K) and the difference in free energies of solvation.

(6)

(7)

where DG0solv,AH is the free energy of solvation of the closed shell neutral species, and

DG0solv,A• is the free energy of solvation of the open shell species, and where ∆G0

H• is

the free energy of solvation of atomic hydrogen. A value of 27.8 kJ/mol was assigned

to ∆G0H based upon the free energy of solvation of H2.8-11

Aqueous Eox values were computed for all anilines investigated in this study.

BDFE values were computed for all phenols considered in this study. These values,

listed together with experimental Eox and BDFE data where available, are listed in

Tables A1 and A2, respectively. Gas phase free energies of reaction were calculated

with Gaussian 09 using CBS-QB3,12,13 or ROCBS-QB314 for the open shell cases

using the default geometry optimization (B3LYP15/6-311-G(2d,d,p)). For thermal

contributions at 298 K, the CBS-QB3 default B3LYP/6-311-G(2d,d,p) thermal free

energy was applied. In the case of the single electron oxidation reactions, the

!

!

AHaq " Aaq• +Haq

!

!

"Gox0 = "Ggas, 298K +""Gsolv

!

!

""Gsolv = "G0solv,A• +"G0

solv,H• # "G0solv,AH

Page 78: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  73  

integrated heat capacity of the electron (0.752 kcal/mol) was also included.16

For the free energies of reaction in implicit solvent, gas phase geometries were

optimized using M06-2X17/6-31+G(d)18,19 with the Gaussian 09 (v. B01) software

package.20 Stationary structures were confirmed by frequency analysis at the same

level of theory. Single point energies using the default cavity were computed with

Gaussian 09 and with the SMD solvation model, and using the MPW1K21/aug-cc-

pVTZ22 electronic structure. As SMD normally employs a standard state of 1 mol/L in

each the gas phase and the aqueous phase, to convert to the standard conditions

described above (with the standard state of 1 atm in the gas phase and 1 mol/L in the

solution phase), 1.89 kcal/mol was added to each computed solvation free energy,

which represents the free energy associated with compressing 1 mol/L to 1 atm.23 As

the solubility of H2 was reported at 1 atm standard state at 298 K, the 1.89 kcal/mol

standard state correction was not applied to the free energy of solvation of H• since its

standard state was already consistent. This means that effectively, this correction ends

up canceling out in both equations 3 and 7 for determining the ∆∆Gsolv for computing

Eox and BDFE, respectively.

In some cases, multiple conformational isomers were found to be stable. In such

cases, the free energy of reaction of conformational isomers were averaged according

to their respective energy of reaction. Thus, the average energy was determined from

the formula:

(8)

Where ∆Grxn is the averaged free energy of reaction (either of Eox or BDFE), R is the

universal gas constant, T is temperature (298 K), and ∆Gi is the free energy of

reaction for each specific conformer i. These included 3-ethoxyaniline, 3-ethylaniline,

!

!

"Grxn = RT ln #i e"Gi RT( )

Page 79: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 74  

3-hydroxyaniline, 3-methoxyaniline, 1-naphthol, 2-naphthol, 2,6-dimethoxyphenol, 3-

aminophenol, 4-ethoxyphenol, 4-ethylphenol, 4-methoxyphenol, 4-acetylphenol,

hydroquinone, and methylhydroquinone.

Results and Discussion

Computation of Aqueous Oxidation Potentials

The complete computed results are listed in Tables A1 and A2. For Eox, the

MAD (mean average deviation) compared to experiment was 0.14 V for the absolute

computation of oxidation potential. This level of accuracy is consistent with a recent

assessment that found that absolute oxidation potentials computed with the SMD

solvation model can have deviations as large as 0.3 V.24 However, computed redox

potentials are often found to correlate well with experimental data within a compound

family.25-28 Hence, a linear regression of computed and experimental Eox data may be

used to improve computed estimates of Eox for compounds where experimental data

are not available. For the set of anilines that we studied, a linear regression of

experimental and computed oxidation potentials produced the relationship:

Eox,fitted = 1.2889 Eox,comp - 0.1584 (r2 = 0.9470) (9)

where Eox,comp are the computed oxidation potential values (in V) and Eox,fitted

represents the best fit to the experimental data. Equation 9 was used to estimate Eox

values of compounds lacking known experimental values, analogous to correlations

previously applied.25-28 Resulting Eox estimates are considered to have an uncertainty

of < 0.1 V, based on the observed root mean squared error of 0.05 V in equation 9.

Page 80: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  75  

Computation of Aqueous bond dissociation free energies

Of the compounds within our data set, only phenol has a reported

experimental gas phase BDFE.11 The computed gas phase BDFE was 79.06 kcal/mol,

only 0.74 kcal/mol below the experimental value of 79.80 kcal/mol. Unfortunately,

one gas phase data point is not enough to be able to evaluate the effectiveness of our

chosen method in the gas phase, (RO)CBS-QB3 for its ability to compute a gas phase

BDFE. In solution, as seen in Table A2, our computational method consistently

underestimated the BEFE values, from anywhere between 2.9 and 4.5 kcal/mol

(except for the hydroquinones). However, the computed BDFE values do, for the

most part, capture the relative difference in BDFE between compounds compared to

phenol. Except for the case of the hydroquinones, this difference in BDFE captured

the relative difference in BDFE from phenol with a MAD of 0.54 kcal/mol,

suggesting that our computational method adequately captures the trends in BDFE

from the substituted groups on the phenols. Thus, in the same manor as was done for

the computed Eox values, the MAD of computed aqueous BDFEs compared to

experiment was 2.81 kcal/mol, and the fitted linear regression was:

BDFEfitted = 0.6245 BDFEcomp + 29.483 (r2 = 0.7494) (10)

Where BDFEcomp are the computed BDFE values (in kcal/mol) and BDFEfitted

represents the best fit to the experimental data. However, the hydroquinones

(hydroquinone, methylhydroquinone) exhibited a rather strong deviation from the

other phenols, and so were excluded from the fitting. The reasons for such deviations

could be due to having two –OH groups that could participate in the hydrogen atom

transfer, the implicit solvation model SMD could not appropriately take into account

Page 81: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 76  

the specific H-bonding interactions for these molecules, or there could be other issues

with the electronic structure method. Removing these values resulted in a fitted linear

regression of:

BDFEfitted = 1.049 BDFEcomp – 7.7751 (r2 = 0.9453) (11)

resulting in a MAD of 1.38 kcal/mol. Equation 11 was used to estimate BDFE values

of compounds lacking known experimental values.

Transient absorption of methylene blue

The majority of our experiments were performed with MB as the triplet state

electron and hydrogen atom acceptor. Compared to many other triplet sensitizers used

in aqueous oxidation experiments, MB has a low triplet energy (138 kJ/mol) and one-

electron reduction potential (-0.23 V vs. NHE), making it a relatively weak oxidant.

Because of the low oxidation potentials of the substituted anilines studied (Eox =

0.53V-1.15V), stronger oxidants react with all of the test compounds at diffusion-

controlled rates. Thus, in order to observe a change in electron transfer rate constants

for the different anilines, a weak oxidant like MB was required. Also, MB is basic

enough to participate in hydrogen atom transfer reactions, and has been used in other

experiments29 for this purpose.

After excitation, methylene blue exhibits three strong features in its transient

absorption spectrum. Two broad positive ∆A signals centered on 420 nm and 830 nm

are both attributed to triplet-triplet absorption, and a strong, negative ∆A signal is

observed at 665 nm, which is the bleaching of the singlet ground state absorption.

With no quenchers present in samples purged with N2, the decays of both triplet

absorptions and the return singlet ground state absorption could be fit with simple

first-order kinetics and a triplet lifetime (t) of 42 µs was observed. In samples purged

with synthetic air, O2 is the only non-solvent quencher present in the system, the

Page 82: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  77  

triplet absorptions at 420 nm and 830 nm both have the same lifetime (1.6 µs) as the

repopulation of the singlet ground state. This gives a second-order rate constant for

MB triplet relaxation by energy transfer to O2 of about 2.4x109 M-1s-1, which is

consistent with other reported rate constants for triplet energy transfer to O2 in MeCN

(1.7x109)30 and MeOH (2.2x109)31. Additionally, no other transient signals are seen.

No power dependance on either the 3MB* lifetime or calculated quenching rate

constants was observed for pulse energies ranging from 2 to 4 µJ. Taken together,

these observations indicate that intersystem crossing (ISC) back to the singlet ground

state and energy transfer to O2 are the only important triplet state loss mechanisms,

when no additional quenchers are present.

Quenching with amine donors

The first series of triplet quenchers studied were substituted aromatic amines.

With amines present in the MB solutions, the lifetime of the 3MB* signal as observed

at 830 nm decreased linearly with amine concentration. In addition to decreasing the

3MB* lifetime, the presence of donors led to the formation of two longer-lived signals

(seen in Figure 2), one of which appeared around 420 nm, thus obscuring the 3MB*

absorbance at this wavelength, and the other around 890 nm. The two additional

signals described appear on different timescales, with the one at 420 nm appearing to

grow directly from the loss of the triplet (this can only be inferred due to the

interfering 3MB* absorption) while the absorption at 890 nm appeared with a delay.

The intensity of the newly observed signals was proportional to the added quencher

concentration. Both additional signals have been previously reported,32,33 with the

absorption at 420 nm coming from the methylene blue neutral radical MB! and the

absorption at 890 nm belonging to the protonated radical cation (MBH!+). The

Page 83: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 78  

quenching behavior and observed transient reaction products provide strong evidence

a one-electron transfer reaction from the amine quenchers to MB.

Figure 1- Structures of MB and other observed transient species.

Figure 2- Selected 3D transient absorption spectra of MB illustrating the spectroscopic differences between electron transfer and PCET quenching. Spectrum A shows the decay of 3MB in an air purged solution. Spectrum B shows MB when quenched by 1.6mM 1,4 benzene diamine. Spectrum C shows MB when quenched by 1.6mM hydroquinone.

N

NN S

N

NN S

N

NN S

+ e- + H

H

3MB+MB MBH+

Page 84: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  79  

The products expected from this reaction would be MB! and an amine radical

cation (RNH2!+). While MB! is clearly visible, RNH2

!+ is not observed. This is most

likely due to the much larger extinction coefficient (e) of the MB species. The

appearance of MBH!+ is explained by the protonation of MB!, which would be

expected to occur at our experimental pH of 8.4 as MBH!+ has a pKa of ~932. The

second-order rate constants (kq) for the reaction of anilines with 3MB* was

determined as the slope of the observed triplet decay rate as a function of aniline

concentration, using the relationship:

(12)

where kobs is the first order decay rate constant of 3MB as calculated from the decay

lifetime where kobs = 1/t, and ko is the decay rate in the absence of quencher. In all

cases, quencher concentrations (0.1-1.6 mM) are in great enough excess to assume

pseudo-first-order behavior, as the concentration of triplet-state MB is no greater than

1 µM. This concentration is estimated from the absorbance of 3MB* and its published

e at 830 nm using Beers law. Table 1 lists the observed quenching rate constants for

the investigated amines. As would be expected for an electron transfer reaction, the

observed second-order reaction rate constants increase with decreasing donor

oxidation potential, reaching a maximum value around 4.5x109M-1s-1 that corresponds

to the diffusion-limited second-order reaction rate constant in water (kd). This value is

kobs = k0 + kq[Q]

Page 85: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 80  

in good agreement with previous reports2,34 of electron transfer reactions in aqueous

solution.

Reaction mechanism and modeling the quenching rate constants of anilines

For the quenching of 3MB* by substituted anilines, we propose the following

mechanism based on the work by Farid et al.35,36:

(13)

In the first step, the free 3MB* and aniline diffuse together in solution to form an

encounter complex. In this rapid step, the reactants are in contact but are only loosely

bound and do not yet have the proper conformation for electron transfer. From this

contact pair, the molecules may either diffuse apart or find the proper orientation for

electron transfer and form an exciplex (shown in brackets). In this longer-lived

species, electron transfer in both the forward and reverse direction (not shown)

occurs. Following electron transfer, the exciplex conformation can be lost and

eventually the pair can diffuse apart. In our system, the evidence for an exciplex is

less direct than for systems where exciplex fluorescence can be directly observed35,36.

Our evidence comes instead from the observed effects on the rate of ISC in the

exciplex seen for MB37 and other triplet dyes38. For anilines containing a heavy atom

substituent such as Br or I, the forward-reverse electron transfer occurring in the

3MB+* + PhNH2

kd

k-d

3MB+* / PhNH2

kc

k-c

3MB•---PhNH2•+

k'-c k'c

3MB• / PhNH2•+ksep

3MB• + PhNH2•+

kISC

1MB+---PhNH2

Page 86: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  81  

exciplex can end in deactivation of 3MB* back to its ground state via ISC without

yielding any apparent reaction. In other words, some anilines containing heavy atoms

can efficiently quench 3MB* while yielding very little MBH!+. This process has been

described previously38 for the triplet quenching of thionine in methanol, where it was

shown that the magnitude of the effect depends on the halogen substitution position

and thus degree of interaction with the ring p system. This influence on the ISC rate

appears to come from the exciplex and as a result of an electron transfer because the

overall rate of quenching kq still depends on the oxidation potential of the aniline. In

other words, the enhancement in ISC is only seen when electron transfer is also

thermodynamically possible.

In a photoinduced electron transfer39-41, the free energy (∆Get) of the reaction

depends on the difference in the electrochemical potentials of the donor and acceptor

plus the excited state energy of the acceptor (or donor):

(14)

where F is Faradays constant and

E 3Sens is the energy of the first excited triplet state

relative to that of the singlet ground state. Based on the theories of electron transfer

reactions proposed by Marcus39, the Rehm-Weller equation has found wide use

because it correctly models the apparent lack of an inverted region, or a slowing of

electron transfer rate for highly exergonic reactions, that is not observed for

photoinduced electron transfers between freely diffusing molecules in solution. A

simplified Rehm-Weller equation is shown below:

ΔGet = −F(Ered − Eox )+ E 3Sens

Page 87: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 82  

kq =kd

1+kdKdZ

exp ΔGet

2+λ4⎛

⎝ ⎜

⎠ ⎟ 2

+ΔGet

2

⎝ ⎜ ⎜

⎠ ⎟ ⎟ /RT

⎢ ⎢

⎥ ⎥ + exp ΔGet

2⎛

⎝ ⎜

⎠ ⎟

⎧ ⎨ ⎪

⎩ ⎪

⎫ ⎬ ⎪

⎭ ⎪

(15)

where Kd = kd/k-d, which represents the equilibrium constant for the formation of the

encounter complex, Z is the universal collision frequency factor, R is the universal

gas constant, and l is known as the solvent reorganization energy. l refers to the

energy required for the system, both the newly formed radical pair and the solvent

cage surrounding it, to rearrange to accommodate the change in charge after the

electron transfer has occurred. This equation correctly models a plateau in the

observed electron transfer rate for highly exergonic reactions (when ∆Get > l) where

kobs = kd and a gradual drop in the rate of electron transfer for exergonic reactions.

The steepness of the drop in rate is affected by the term l. The Rehm-Weller equation

has been used extensively to semi-empirically model photoinduced electron transfer

reactions in a wide range of systems, however a simplified alternative explanation for

the behavior observed by Rehm and Weller has been put forth by Farid and

coworkers35,36 arguing that for some systems electron transfer reaction rate constants

can be modelled by a simple Sandros-Boltzmann distribution:

(16)

where s is a positive free energy shift associated with the difference in free energy of

the radical pair in contact relative to that of the free radicals in solution. A similar

shift in free energy was used in the original Rehm-Weller description of the reaction

free energy, however in their case it was argued that a negative shift in free energy

was required. At the time it was not established that the free energy of a radical pair is

higher than that of the free radicals35 for these reactions.

!

!

kobs =kd

1+ exp "Get + sRT

#

$ %

&

' (

Page 88: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  83  

Figure 3 shows a plot of log kq for the observed quenching rate constants

between 3MB* and the substituted anilines vs. ∆Get as calculated by equation 14 using

both experimental and computationally derived Eox values. For anilines that had

reported Eox values determined by pulse radiolysis ∆Get The rate constant for electron

transfer increases with decreasing ∆Get, until reaching a plateau at the diffusion

controlled limit so that kq = kd. For the quenching of 3MB* by anilines, we observed

that kq begins to fall below kd at relatively low values of ∆Get, around -16 kJ/mol.

Going toward larger ∆Get values, kq decreases with a slope of approximately 1/RT.

For these data, the best fit was achieved using the Sandros-Boltzmann equation where

kd= 4.3x109 M-1s-1 and s = 16.2 kJ/mol. The original work by Farid et al.35 describing

the use of the Sandros-Boltzmann equation to fit electron transfer data reported a s of

7.7 kJ/mol for electron transfer reactions between singlet excited pyrylium cation

acceptors and neutral aromatic hydrocarbon donors in acetonitrile.

Page 89: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 84  

Figure 3- Plot of ∆Get vs. kq for the electron transfer reaction from substituted anilines to MB. ∆Get was calculated from computational Eox values (red triangles) and pulse radiolysis values (blue triangles) the lines depict the Sandros-Boltzmann fits for the computational (solid) and pulse radiolysis (dashed) ∆Get values. The error bars on the computed values indicate the estimated 0.1 V error in the calculations.

The reason that we observe a larger s value compared to the previous work is

not immediately clear. In both cases, the acceptor is initially positively charged, and

following electron transfer becomes neutral, there are important differences, however,

in the charge delocalization before and after electron transfer in each system. In the

pyrylium case, the initial positive charge is localized entirely on the central pyrylium

ring and transferred to a single aromatic ring following electron transfer. In our case,

the positive charge on MB is delocalized over all three fused rings in the molecule

before electron transfer, but then transferred to a single ring on an aniline. This

difference in charge density could contribute to the additional energy requirement

seen in electron transfers to MB. The solvent (water vs. acetonitrile) is another factor

that might add to the increased s in our system, but predicting its effect is less

straightforward than for the role of charge density. Lastly, there is the difference in

spin states of the acceptors, as we are following the quenching of a triplet state rather

-80 -70 -60 -50 -40 -30 -20 -10 0

107

108

109

1010

k q (M-1s-1

)

ΔG (kJ/mol)

Page 90: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  85  

than a singlet. There exists conflicting results as to weather spin multiplicity affects

the observed kq relative to ∆Get, with some arguing there are no differences42, and

others reporting37,43,44 lower kq values for triplet states at equivalent ∆Get.

Table 2- Data for substituted anilines at pH 8.4

Table 3- Data for substituted phenols at pH 7.0

a: BDFEcalc values were calculated from the phenol pKa and PhO- Eox values according to the procedure described in ref. 11. b: BDFEcomp values were computed as described in the methods.

Substituted aniline Computational Eox Pulse radiolysis Eox kq (108 M-1s-1)4-N,N-dimethylamino aniline 0.39 48.5 ± 1.64-methoxyaniline 0.82 0.79 45.8 ± 1.1N-methylaniline 0.90 43.0 ± 1.74-aminoaniline 0.54 0.59 42.4 ± 1.34-methylaniline 0.95 0.92 41.5 ± 0.54-tert-butylaniline 0.98 0.95 32.7 ± 0.93-hydroxyaniline 1.08 18.3 ± 0.143-ethylaniline 1.10 15.4 ±0.43-methylaniline 1.02 15.2 ± 0.73-ethoxyaniline 1.09 13.0 ± 0.33-methoxyaniline 1.09 11.7 ± 0.44-fluoroaniline 1.07 8.65 ± 0.16aniline 1.07 1.02 4.10 ± 0.054-chloroaniline 1.12 1.01 3.79 ± 0.163-fluoroaniline 1.18 0.109 ± 0.013-chloroaniline 1.20 0.106 ± 0.01

Substituted phenol pKa PhO- EoxaBDFEcalc

bBDFEcomp kq (108 M-1s-1)sesamol 9.7 0.45 81.3 81.8 39.7 ± 0.4methylhydroquinone 10.1 0.4 80.7 83.8 39.1 ± 0.8hydroquinone 9.85 0.45 81.5 84.7 24.1 ± 0.31-naphthol 9.3 0.59 83.9 83.3 16.8 ± 0.42,6-dimethoxyphenol 9.98 0.47 82.1 84.6 16.2 ± 0.54-ethoxyphenol 10.2 0.52 83.6 84.4 12.2 ± 0.54-methoxyphenol 10.1 0.54 83.9 84.5 7.93 ± 0.042-naphthol 9.6 0.69 86.7 86.9 3.66 ± 0.34-chlorophenol 9.4 0.8 88.9 89.4 0.96 ± 0.074-ethylphenol 10 0.68 87.0 87.1 0.66 ± 0.194-fluorophenol 9.9 0.76 88.7 88.4 0.63 ± 0.074-methylphenol 10.3 0.68 87.4 87.1 0.57 ± 0.094-tertbutylphenol 10.23 0.76 89.1 87.5 0.52 ± 0.03phenol 10 0.79 89.5 89.4 0.17 ± 0.03

Page 91: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 86  

We also attempted to explain the data using the Rehm-Weller equation. Reasonable

fits were only obtained after introducing a correction to the free energy:

(17)

This correction has been used before to address uncertainties in donor/acceptor redox

values2, and also to address shifts in ∆Get that seem to be system-specific37. Using

adjusted ∆Get values gave reasonable fits, but with very small values for l that are

typically seen in much more rigid systems.45,46 For this reason, we favor the fits given

by the Sandros-Botzmann equation for the purposes of predicting kq rates for electron

transfer reactions between MB and anilines.

Quenching with phenol donors

The second class of quenchers we investigated was substituted phenols. The

quenching rate constants for phenols were again calculated using equation 12. The

results at pH 7 are shown in table 2. Phenolic quenching of 3MB* yielded observable

reaction products in the transient absorption spectra that were indicative of the

reaction mechanism. The addition of phenols resulted in a rapid appearance of MBH!+

(Figure 2) with no lag as seen with anilines. For strong quenchers, the conversion

from 3MB* to MBH!+ was nearly quantitative based on their respective loss/formation

rates at high phenol concentration. In the region of 410-430 nm the absorption is

initially dominated by the 3MB* signal, but due to quenching of this species, a distinct

change in the absorption shape and lmax at longer times is observed. We attribute this

absorption change to the formation of phenoxyl radicals (PhO!), as the new lmax agrees

with reported values for previously reported phenoxyl radicals47.

ΔGet = ΔGet +δΔGet

Page 92: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  87  

Phenolic quenching mechanism

This observed behavior is explained best by a proton-coupled electron transfer

(PCET) quenching mechanism, where rather than donating only an electron, phenols

also transfer a proton in the same or a rapidly sequential step. Another term for this

mechanism is hydrogen atom transfer, however we favor the assignment of PCET

because, in a strict sense hydrogen atom transfer requires that the electron come

directly from the same bonding orbital that the proton was associated with, while in

PCET the electron may come instead from an orthogonal p orbital11. The former is

often the case for phenols11,48, and also likely in our system. PCET from phenols is a

well-known phenomenon47-51 that is widely observed in natural systems, and is partly

responsible for the anti-oxidant role of phenols in biological systems.

In addition to the observed spectroscopic evidence for PCET, the

thermodynamics of the process are strongly against the plausibility of a simple one-

electron transfer from phenol to 3MB*. For example, phenol has a one-electron

oxidation potential of 1.5 V, so according to equation 14, the free energy of 3MB*

quenching by electron transfer from phenol would be 28.9 kJ/mol, which by

calculation with equations 15 or 16 would yield a kq several orders of magnitude

smaller than what is observed in our experiments. Our proposed reaction mechanism

for PCET from phenols to 3MB* is shown in equation 17.

(17)

In the first step, 3MB* and PhOH form an exciplex. Within this exciplex, PhOH

donates both an electron and a proton in a sequential manor for a net hydrogen atom

3MB+* + PhOHk1

k-1

3MB+*---HOPhk2 MBH+ ---OPh MBH+ + OPh

k3

Page 93: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 88  

transfer. Finally the individual radicals are free to diffuse away from one another in

solution.

To further test the hypothesized mechanism, we repeated the phenolic

quenching experiments in D2O. Due to rapid exchange of the protons between the

phenolic OH group and solvent, nearly all of the phenol under this condition can be

considered deuterated. Under these conditions, we would expect PCET from a

deuterated phenol and 3MB* to occur with a reduced rate constant due to the

increased OD bond strength. Figure 4 shows an example of these experiments. As

expected, a reduction in the quenching rate constants were observed for each phenol

tested. H/D isotope effects, as determined by kH/kD, ranged from 1.6 to 3.3, with the

low values only coming from quenchers reacting at or near the diffusion controlled

limit.

Figure 4- Observed decay rates of 3MB in the presence of 4-methoxyphenol at pH/D= 7. The red points were collected in H2O and the blue in D2O.

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.00.5

1.0

1.5

2.0

2.5

k obs (s

-1) x

106

[4-methoxyphenol] (mM)

Page 94: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  89  

These values are in good agreement with what has been previously observed

for phenolic quenching of aromatic ketones by PCET50 as well as for intramolecular

PCET reactions48. Alternatively, when experiments were done at high pH (or pD),

ensuring only phenolate being present, quenching by electron transfer took place and

no kinetic isotope effect was observed. In taken together, both transient absorption

spectroscopy data and kinetic isotope experiment data are consistent with phenolic

quenching of 3MB* via a PCET mechanism, not an electron transfer mechanism.

In equation 17, the rate-determining step is the breaking of the phenolic OH

bond. It follows from this that a correlation between the OH bond dissociation free

energy (BDFE) would be expected in the 3MB* quenching rate constants for phenols.

A plot of BDFE vs. kq for the phenolic quenchers is shown in Figure 5. In addition to

the computationally derived BDFE values discussed earlier, BDFEs were also

calculated11 from a simple thermodynamic cycle based on a two-step reaction where

an electron and a proton are lost. This method of calculation requires only the

phenolate Eox, pKa and a term for the solvation of a hydrogen atom CG, all of which

are solvent-dependent. As seen in the Figure 5, quenching rate constants correlate

strongly with BDFE. Notable, however, is the systematic nature of the deviation seen

in the plot. The points in Figure 5 are color mapped according to the phenols’ pKa

values. A clear trend is seen in which the more acidic phenols react with an enhanced

rate when compared to a more basic phenol with a similar BDFE. The origin of this

acidity enhancement may come in the mechanistic step that precedes PCET. Before

PCET can occur, 3MB* must first form a hydrogen-bonded exciplex with the phenol.

More acidic phenols carry a stronger ∂+ on the OH hydrogen, which has the known

effect of increasing hydrogen-bonding strength.52 This would lead to a stronger

complex between more acidic phenols and 3MB*. In reaction equation 17, the

Page 95: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 90  

equilibrium constant K1=k1/k-1 would be proportional to the phenol pKa, with more

acidic phenols enhancing the reaction K1, thus enhancing the observed rate of PCET.

Another possible explanation of this effect would be the greater amount of the

ionized phenolate form present at the experimental pH of 7, which is a more efficient

quencher (via electron transfer) than the protonated parent phenol. This possibility is

ruled out based on the very low amount of phenolate (less than 0.5% ionized) for even

the most acidic phenols studied, which would add an error of less than 2%.

Figure 5- Plot of kq vs. substituted phenol BDFE for the reaction between substituted phenols and 3MB The best fit line is shown.

80 84 88 92

k q((M

+1s+1 )

BDFE((kcal/mol)

9.0

9.5

10.0

10.5

(

pKa

107

108

109

1010

Page 96: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  91  

Figure 6- Plot of kq vs. ∆GPCET for the reaction between substituted phenols and 3MB. The best fit line is shown.

Modeling kq for phenols

Finally, we attempted to find a linear free energy relationship capable of

predicting quenching rate constants of 3MB* by substituted phenols. As already

demonstrated to some degree in by the D2O experiments, the reaction rate constant is

primarily determined by the phenolic O-H bond strength. The result of this is that the

difference in free energy of the MB-H bond formed and the PhO-H bond broken

should show a linear correlation with the observed quenching rate of 3MB*. Therefore

it should be possible to use the change in free energy (∆GPCET) from the BDFEs

before and after PCET using equation 18 and determine the empirical relationship

between ∆GHAT and kq.

ΔGPCET = BDFE(PhO ⋅ ⋅ ⋅H )− BDFE(MB ⋅ ⋅ ⋅H ) (18)

!75 !70 !65 !60 !55 !50 !45 !40 !35 !30

107

108

109

1010

+

+

k q+(M

!1s!1 )

�GPCET

+(kJ/mol)

9.5

9.8

10.1

10.4

pKa

Page 97: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 92  

Figure 6 shows ∆GPCET plotted against kq. The result is once again linear, and still

shows a deviation with pKa. An empirical equation predicting kq is obtained by fitting

both ∆GPCET and pKa with a non-linear regression, the results of which are shown in

Figure 7. This fit achieves good agreement between predicted and observed kq values

for phenolic quenching of 3MB*.

Figure 7- Plot of observed vs predicted kq values for PCET from substituted phenols to 3MB. The blue line indicates a 1:1 agreement.

Implications and Conclusions

In this work, we demonstrate that for a single photoexcited oxidant, 3MB*, the

reaction rate constant and mechanism can change depending on the oxidation

107 108 109 1010107

108

109

1010

Obs

erve

d k q (M

-1s-1

)

Predicted kq (M-1s-1)

Page 98: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  93  

potentials and BDFEs of potential donors in aqueous systems. The development of

empirical models capable of predicting reaction rate constants is a key part in

determining the environmental persistence of organic pollutants, and this work was

intended to aid in this area. For anilines it is clear that in order to predict kq values, a

reliable method for determining Eox is first required to obtain ∆Get. The irreversible

nature of the one-electron oxidation for anilines and other compounds of interest

mean that high quality Eox values are often unavailable and difficult to obtain

experimentally. Furthermore, many pollutants have low water solubilities that

complicate aqueous Eox determination. In this work, we explored the use of

computational methods to bridge this gap. However, it is also clear that for electron

transfer reactions where ket is quite sensitive to small changes in reactant Eox values,

small errors in computational methods can yield large errors in rate constant

prediction. Further work is required to improve these computational methods if more

reliable predictions are to be made.

One aspect of this work that can be improved in future studies with respect to

the environmental relevance is the identity of the photooxidant, MB. In aquatic

environments, triplet-state dissolved organic matter (3DOM) is one of the main drivers

of photoinduced oxidation for pollutants, and the main components of DOM

responsible for the oxidative properties are believed to be aromatic ketones53. Similar

comparative studies using anilines and phenols with aromatic ketone sensitizers are a

clear avenue for further studies. While the chemistry of MB is distinct from aromatic

ketone photosensitizers, it may be a reasonable model for other environmentally

relevant compounds, namely flavins. Future work is required to test the broader

applicability of our results to other systems and importantly, more environmentally

relevant oxidants.

Page 99: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 94  

References:

(1) Canonica, S.; Hellrung, B.; Müller, P.; Wirz, J. Aqueous Oxidation of Phenylurea Herbicides by Triplet Aromatic Ketones. Environmental Science & Technology 2006, 40, 6636-6641. (2) Canonica, S.; Hellrung, B.; Wirz, J. Oxidation of Phenols by Triplet Aromatic Ketones in Aqueous Solution. The Journal of Physical Chemistry A 2000, 104, 1226-1232. (3) Canonica, S.; Jans, U.; Stemmler, K.; Hoigne, J. Transformation Kinetics of Phenols in Water: Photosensitization by Dissolved Natural Organic Material and Aromatic Ketones. Environmental Science & Technology 1995, 29, 1822-1831. (4) Werner, J. J.; McNeill, K.; Arnold, W. A. Environmental photodegradation of mefenamic acid. Chemosphere 2005, 58, 1339-1346. (5) Wenk, J.; Eustis, S. N.; McNeill, K.; Canonica, S. Quenching of Excited Triplet States by Dissolved Natural Organic Matter. Environmental Science & Technology 2013, 47, 12802-12810. (6) Mohr, P. J.; Taylor, B. N.; Newell, D. B. 2011; Vol. 2013. (7) Truhlar, D. G.; Cramer, C. J.; Lewis, A.; Bumpus, J. A. J. Chem. Educ. 2004, 81, 596–604. Journal of Chemical Education 2007, 84, 934. (8) In IUPAC Solubiligy Data Series; Vol. 5/6 ed.; Young, C. L., Ed.; Pergamon: New York, 1981. (9) Roduner, E. Hydrophobic solvation, quantum nature, and diffusion of atomic hydrogen in liquid water. Radiat. Phys. Chem. 2005, 72, 201-206. (10) Stepanic, V.; Troselj, K. G.; Lucic, B.; Markovic, Z.; Amic, D. Bond dissociation free energy as a general parameter for flavonoid radical scavenging activity. Food Chem., 141, 1562-1570. (11) Warren, J. J.; Tronic, T. A.; Mayer, J. M. Thermochemistry of Proton-Coupled Electron Transfer Reagents and its Implications. Chemical Reviews 2010, 110, 6961-7001. (12) Montgomery, J. A.; Frisch, M. J.; Ochterski, J. W.; Petersson, G. A. A complete basis set model chemistry. VI. Use of density functional geometries and frequencies. J. Chem. Phys. 1999, 110, 2822-2827. (13) Montgomery, J. A.; Frisch, M. J.; Ochterski, J. W.; Petersson, G. A. A complete basis set model chemistry. VII. Use of the minimum population localization method. J. Chem. Phys. 2000, 112, 6532-6542. (14) Wood, G. P. F.; Radom, L.; Petersson, G. A.; Barnes, E. C.; Frisch, M. J.; Montgomery, J. A. A restricted-open-shell complete-basis-set model chemistry. J. Chem. Phys. 2006, 125, 16. (15) Becke, A. D. DENSITY-FUNCTIONAL THERMOCHEMISTRY .3. THE ROLE OF EXACT EXCHANGE. J. Chem. Phys. 1993, 98, 5648-5652. (16) Bartmess, J. E. THERMODYNAMICS OF THE ELECTRON AND THE PROTON. J. Phys. Chem. 1994, 98, 6420-6424. (17) Zhao, Y.; Truhlar, D. G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215-241. (18) Harihara, P. C.; Pople, J. A. INFLUENCE OF POLARIZATION FUNCTIONS ON MOLECULAR-ORBITAL HYDROGENATION ENERGIES. Theor. Chim. Acta 1973, 28, 213-222.

Page 100: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  95  

(19) Hehre, W. J.; Ditchfie.R; Pople, J. A. SELF-CONSISTENT MOLECULAR-ORBITAL METHODS .12. FURTHER EXTENSIONS OF GAUSSIAN-TYPE BASIS SETS FOR USE IN MOLECULAR-ORBITAL STUDIES OF ORGANIC-MOLECULES. J. Chem. Phys. 1972, 56, 2257-&. (20) Frisch, M. J. T., G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. 2009. (21) Lynch, B. J.; Fast, P. L.; Harris, M.; Truhlar, D. G. Adiabatic connection for kinetics. J. Phys. Chem. A 2000, 104, 4811-4815. (22) Kendall, R. A.; Dunning, T. H.; Harrison, R. J. ELECTRON-AFFINITIES OF THE 1ST-ROW ATOMS REVISITED - SYSTEMATIC BASIS-SETS AND WAVE-FUNCTIONS. J. Chem. Phys. 1992, 96, 6796-6806. (23) Lewis, A.; Bumpus, J. A.; Truhlar, D. G.; Cramer, C. J. Molecular Modeling of environmentally important processes: Reduction potentials. Journal of Chemical Education 2004, 81, 596-604. (24) Guerard, J. J.; Arey, J. S. Critical Evaluation of Implicit Solvent Models for Predicting Aqueous Oxidation Potentials of Neutral Organic Compounds. J. Chem. Theory Comput., 9, 5046-5058. (25) Arnold, W. A. One electron oxidation potential as a predictor of rate constants of N-containing compounds with carbonate radical and triplet excited state organic matter. Environmental Science: Processes & Impacts 2013. (26) Tentscher, P. R.; Eustis, S. N.; McNeill, K.; Arey, J. S. Aqueous Oxidation of Sulfonamide Antibiotics: Aromatic Nucleophilic Substitution of an Aniline Radical Cation. Chem.-Eur. J., 19, 11216-11223. (27) Winget, P.; Cramer, C. J.; Truhlar, D. G. Computation of equilibrium oxidation and reduction potentials for reversible and dissociative electron-transfer reactions in solution. Theor. Chem. Acc. 2004, 112, 217-227. (28) Winget, P.; Weber, E. J.; Cramer, C. J.; Truhlar, D. G. Computational electrochemistry: aqueous one-electron oxidation potentials for substituted anilines. Physical Chemistry Chemical Physics 2000, 2, 1231-1239. (29) Atherton, S. J.; Harriman, A. Photochemistry of intercalated methylene blue: photoinduced hydrogen atom abstraction from guanine and adenine. Journal of the American Chemical Society 1993, 115, 1816-1822. (30) Montalti, M.; Credi, A.; Prodi, L.; Gandolfi, M. T. Handbook of photochemistry; CRC press, 2006. (31) Rehak, V.; Poskocil, J. LASER FLASH-PHOTOLYSIS OF METHYLENE-BLUE SOLUTIONS. Collect. Czech. Chem. Commun. 1979, 44, 2015-2023. (32) Ohno, T.; Lichtin, N. N. Electron transfer in the quenching of triplet methylene blue by complexes of iron(II). Journal of the American Chemical Society 1980, 102, 4636-4643.

Page 101: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 96  

(33) Kayser, R. H.; Young, R. H. THE PHOTOREDUCTION OF METHYLENE BLUE BY AMINES—I. A FLASH PHOTOLYSIS STUDY OF THE REACTION BETWEEN TRIPLET METHYLENE BLUE AND AMINES. Photochemistry and Photobiology 1976, 24, 395-401. (34) Behar, D.; Neta, P.; Schultheisz, C. Reaction Kinetics in Ionic Liquids as Studied by Pulse Radiolysis:   Redox Reactions in the Solvents Methyltributylammonium Bis(trifluoromethylsulfonyl)imide and N-Butylpyridinium Tetrafluoroborate. The Journal of Physical Chemistry A 2002, 106, 3139-3147. (35) Farid, S.; Dinnocenzo, J. P.; Merkel, P. B.; Young, R. H.; Shukla, D. Bimolecular Electron Transfers That Follow a Sandros−Boltzmann Dependence on Free Energy. Journal of the American Chemical Society 2011, 133, 4791-4801. (36) Farid, S.; Dinnocenzo, J. P.; Merkel, P. B.; Young, R. H.; Shukla, D.; Guirado, G. Reexamination of the Rehm–Weller Data Set Reveals Electron Transfer Quenching That Follows a Sandros–Boltzmann Dependence on Free Energy. Journal of the American Chemical Society 2011, 133, 11580-11587. (37) Tamura, S.-I.; Kikuchi, K.; Kokubun, H.; Usui, Y. The Electron Transfer Reactions between Triplet Dyes and Aromatic Compounds in Acetonitrile. Zeitschrift für Physikalische Chemie 1978, 111, 7-18. (38) Steiner, U.; Winter, G. Position dependent heavy atom effect in physical triplet quenching by electron donors. Chemical Physics Letters 1978, 55, 364-368. (39) Marcus, R. A. On the Theory of Oxidation-Reduction Reactions Involving Electron Transfer. I. The Journal of Chemical Physics 1956, 24, 966-978. (40) Rehm, D.; Weller, A. KINETICS AND MECHANICS OF ELECTRON TRANSFER DURING FLUORESCENCE QUENCHING IN ACETONITRILE. Berichte Der Bunsen-Gesellschaft Fur Physikalische Chemie 1969, 73, 834-&. (41) Rehm, D.; Weller, A. Kinetics of fluorescence quenching by electron and H-atom transfer. Isr. J. Chem 1970, 8, 259-272. (42) Neumann, M. G.; Pastre, I. A.; Previtali, C. M. Comparison of photoinduced electron transfer to singlet and triplet states of safranine T. Journal of Photochemistry and Photobiology A: Chemistry 1991, 61, 91-98. (43) Kikuchi, K.; Tamura, S.-I.; Iwanaga, C.; Kokubun, H.; Usui, Y. The Electron Transfer Reaction between Triplet Methylene Blue and Aromatic Compounds. Zeitschrift für Physikalische Chemie 1977, 106, 17-24. (44) Bertolotti, S. G.; Montejano, H. A.; Previtali, C. M. Comparison of the Kinetics of Electron Transfer in the Diffusion Limit for the Singlet and Triplet Quenching of Eosin Y by Quinones. Photochemistry and Photobiology 2013, 89, 1442-1447. (45) Imahori, H.; Tkachenko, N. V.; Vehmanen, V.; Tamaki, K.; Lemmetyinen, H.; Sakata, Y.; Fukuzumi, S. An Extremely Small Reorganization Energy of Electron Transfer in Porphyrin−Fullerene Dyad. The Journal of Physical Chemistry A 2001, 105, 1750-1756. (46) Iwaki, M.; Kumazaki, S.; Yoshihara, K.; Erabi, T.; Itoh, S. Delta G(0) dependence of the electron transfer rate in the photosynthetic reaction center of plant photosystem I: Natural optimization of reaction between chlorophyll a (A(0)) and quinone. J. Phys. Chem. 1996, 100, 10802-10809. (47) Lathioor, E. C.; Leigh, W. J. Bimolecular Hydrogen Abstraction from Phenols by Aromatic Ketone Triplets†. Photochemistry and Photobiology 2006, 82, 291-300.

Page 102: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  97  

(48) Rhile, I. J.; Markle, T. F.; Nagao, H.; DiPasquale, A. G.; Lam, O. P.; Lockwood, M. A.; Rotter, K.; Mayer, J. M. Concerted Proton−Electron Transfer in the Oxidation of Hydrogen-Bonded Phenols. Journal of the American Chemical Society 2006, 128, 6075-6088. (49) Di Meo, F.; Lemaur, V.; Cornil, J.; Lazzaroni, R.; Duroux, J.-L.; Olivier, Y.; Trouillas, P. Free Radical Scavenging by Natural Polyphenols: Atom versus Electron Transfer. The Journal of Physical Chemistry A 2013, 117, 2082-2092. (50) Leigh, W. J.; Lathioor, E. C.; St. Pierre, M. J. Photoinduced Hydrogen Abstraction from Phenols by Aromatic Ketones. A New Mechanism for Hydrogen Abstraction by Carbonyl n,π* and π,π* Triplets. Journal of the American Chemical Society 1996, 118, 12339-12348. (51) Yoshihara, T.; Yamaji, M.; Itoh, T.; Shizuka, H.; Shimokage, T.; Tero-Kubota, S. Hydrogen atom transfer and electron transfer reactions in the triplet [small pi],[small pi]* state of 1,4-anthraquinone studied by CIDEP techniques and laser flash photolysis[dagger]. Physical Chemistry Chemical Physics 2000, 2, 993-1000. (52) Anslyn, E. V.; Dougherty, D. A. Modern Physical Organic Chemistry; University Science, 2006. (53) Canonica, S. Oxidation of Aquatic Organic Contaminants Induced by Excited Triplet States. CHIMIA International Journal for Chemistry 2007, 61, 641-644. (54) Jonsson, M.; Lind, J.; Eriksen, T. E.; Merenyi, G. Redox and Acidity Properties of 4-Substituted Aniline Radical Cations in Water. Journal of the American Chemical Society 1994, 116, 1423-1427. (55) Jonsson, M.; Lind, J.; Merenyi, G.; Eriksen, T. E. REDOX PROPERTIES OF 4-SUBSTITUTED ARYL METHYL CHALCOGENIDES IN WATER. J. Chem. Soc.-Perkin Trans. 2 1995, 67-70. (56) Jonsson, M.; Wayner, D. D. M.; Lusztyk, J. Redox and acidity properties of alkyl- and arylamine radical cations and the corresponding aminyl radicals. J. Phys. Chem. 1996, 100, 17539-17543.

Page 103: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 98  

Appendix to Chapter 2

Compound Eox,expt

a Eox,comp Eox,fitted 3-t-fluoroaniline 1.27 1.50 1.30

3-chloroaniline 1.38 1.20 3-ethoxyaniline 1.22 1.09

3-ethylaniline 1.25 1.10 3-fluoroaniline 1.36 1.18

3-hydroxyaniline 1.22 1.08 3-methoxyaniline 1.23 1.09

3-nitroaniline 1.53 1.31 3-toluidine 1.14 1.02

4-acetylaniline 1.14 1.38 1.20 4-bromoaniline 1.04 1.32 1.15 4-chloroaniline 1.01 1.27 1.12 4-cyanoaniline 1.32 1.53 1.31 4-fluoroaniline 1.20 1.07

4-hydroxyaniline 0.76 0.91 0.85 4-t-butylaniline 0.95 1.08 0.98

4-t-fluoroaniline 1.28 1.50 1.30 4-toluidine 0.92 1.05 0.95

aniline 1.02 1.21 1.07 4-anisidine 0.79 0.87 0.82

benzene-1,4-diamine 0.59 0.51 0.54 N-methylaniline 0.95 0.99 0.90

N,N’-dimethylbenzene-1,4-diamine 0.82 0.39 Table A1- Experimental, computed, and fitted single electron oxidation potentials for substituted anilines in aqueous solution vs. SHE (V). a. Ref54-56

Page 104: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  99  

Compund BDFEexpta BDFErel,expt BDFEcomp BDFErel,comp BDFEfitted

1-naphthol 83.90 -5.60 79.37 -6.66 79.87 2-naphthol 86.70 -2.80 83.64 -2.39 86.72

2,6-dimethoxyphenol 80.07 -5.95 81.01 3-aminophenol 84.86 -1.17 88.67 4-chlorophenol 88.90 -0.60 85.94 -0.09 90.40 4-ethoxyphenol 80.84 -5.19 82.24

4-ethylphenol 85.29 -0.74 89.35 4-fluorophenol 88.70 -0.80 84.83 -1.20 88.62

4-methoxyphenol 83.90 -5.60 81.00 -5.03 82.49 4-methylphenol 87.40 -2.10 83.63 -2.40 86.70 4-acetylphenol 89.21 +3.18 95.63

4-t-butylphenol 83.97 -2.05 87.25 phenol 89.5 0.00 86.03 0.00 90.54

sesamol 77.53 -8.50 76.93 hydroquinone 81.5 -8.00 81.33 -4.70 83.02

methylhydroquinone 80.7 -8.80 81.23 -4.80 82.85 Table A2- Experimental, computed, and fitted bond dissociation free energies for substituted phenols in aqueous solution (kcal/mol). BDFErel values are determined as the difference in BDFE relative to phenol. a. Ref11

Page 105: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 100  

Figure A1- A. Thermodynamic cycle used for the computation of Eox in aqueous solution. B. Thermodynamic cycle used for the computation of BDFE in aqueous solution.

-

-

g

aq

g g

gaq

∆Gsolv,A ∆G ∆G = 0

IE

∆Gox,aq

gas, 298 K

0

00

A

AA

A

g

aq

g g

aqaq

∆Gsolv,A ∆Gsolv

∆G

∆Grxn, aq

rxn,g, 298 K

00

A

AAH

AH

∆Gsolv

H

H

A

B

Page 106: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  101  

Figure A2- Linear regression fit of experimental vs. computed aqueous oxidation potential values (V vs. NHE).

Figure A3- Linear regression fit of experimental vs. computed aqueous BDFE values (kcal/mol)

y  =  1.2889x  -­‐  0.1584  R²  =  0.94697  

0.00  

0.20  

0.40  

0.60  

0.80  

1.00  

1.20  

1.40  

1.60  

1.80  

0.5   0.6   0.7   0.8   0.9   1   1.1   1.2   1.3   1.4   1.5  

Computed  Eox  (V)  

Experiment  Eox  (V)  

y  =  1.049x  -­‐  7.7751  R²  =  0.94532  

79  

80  

81  

82  

83  

84  

85  

86  

87  

83   84   85   86   87   88   89   90  

BDFE,com

puted  (kcal/mol)  

BDFE,experiment  (kcal/mol)  

Page 107: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 102  

Page 108: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  103  

Chapter 4

Investigating the mechanism of the

photochemical chlorination of pyrene in

aqueous solution Paul R. Erickson, Kenshi Sankoda, and Kristopher McNeill

In preperation for Environmental Science and Technology: Letters

Page 109: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 104  

Abstract

The reaction mechanism responsible for the photochemical formation of 1-

chloropyrene (Cl-PYR) from pyrene and chloride in aqueous solution was studied.

Earlier work has demonstrated that the presence of Cl-PYR in river sediments leading

to the ocean strongly correlates with the salinity of the sampling site, strongly

suggesting that Cl-PYR is being generated in situ at these sites. Using a combination

of steady-state and laser flash photolysis, we provide evidence that supports the

hypothesis that Cl-PYR is formed from the nucleophilic attack of chloride on pyrene

radical cations which are formed by the oxidation of excited state pyrene by oxygen.

Addition of electron donors to photolysis solutions greatly decreased the production

of Cl-PYR, and when chloride was put in competition with other nucleophiles, the

product distributions indicated chloride behaves like a nucleophile in this system. All

of our results taken together suggest this could be a general mechanism for the

formation of chlorinated polycyclic aromatic hydrocarbons found in marine

environments.

Page 110: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  105  

Introduction

Polycyclic aromatic hydrocarbons (PAHs) are a large class of compounds

comprised of fused aromatic ring structures of varying size and shape, containing only

carbon and hydrogen. PAHs are well-known pollutants, which enter the environment

as either byproducts of incomplete combustion (pyrogenic) or from oil sources

(petrogenic). They are also used in the commercial production of dyes and other

products. The concern regarding PAHs derives from the toxic properties of some

structures, which have been found to be carcinogenic1,2, mutagenic3 as well as

teratogenic4,5 to a wide range of organisms. Their toxicity has led the U.S.

Environmental Protection Agency to label 16 congeners as “priority pollutants”. Once

in the environment, PAHs may transform via a wide range of biotic6 and abiotic7

pathways, sometimes yielding transformation products that are also of concern. One

such example is the chlorination of PAHs. A number of chlorinated PAH (Cl-PAH)

congeners have been found in air3,8-10, biota11, and sediment samples12,13. Cl-PAHs

have similar toxic effects as their parent PAHs, and in some cases have been reported

to exhibit higher toxicity. In particular, Cl-PAHs show higher affinity to aryl

hydrocarbon receptors14, which is one of the main routes of toxicity for dibenzo-p-

dioxins and related compounds.

The chlorination of PAHs can occur during combustion, water chlorination15-

17, and photochemically in the presence of chloride18. The importance of each of these

formation routes to the total environmental concentration present will likely differ by

location, but for marine environments the largest source may be photochemical

1-Chloropyrene

Cl

Pyrene

Page 111: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 106  

generation. Sankoda and coworkers recently reported13 finding Cl-PAHs in river and

tidal flat sediments near the Yabe river in Japan, while Cl-PAHs were not detected in

nearby samples of potential PAH sources such as road dust. Furthermore, the

concentrations of Cl-PAHs correlated positively with the salinity of the sampling

sites. Sankoda, et al. also performed laboratory experiments showing that PAHs in

artificial seawater yield Cl-PAHs when irradiated with light of wavelengths between

300-400 nm. This work, along with other reports,18-20 lends evidence to the hypothesis

that Cl-PAHs found in marine environments are, to a large degree, generated

photochemically. While several studies have hinted at the possible reaction

mechanism of photochemical Cl-PAH formation, a systematic investigation has yet to

be done. The goal of this work was to identify the reaction mechanism by which

PAHs are photochlorinated. We specifically examined pyrene (PYR) and its

transformation to 1-chloropyrene (Cl-PYR) in aqueous solution. We present a series

of experiments testing our hypothesis that Cl-PYR is formed by nucleophilic attack of

chloride on PYR radical cations (PYR!+), which are in turn generated by the oxidation

of excited singlet state PYR by O2 in aqueous solution.

Methods

Steady-state photolysis experiments were performed in a Rayonet photoreactor

containing six 365 nm bulbs. Samples (10 mL) of PYR were made in borate buffer

(9.5 mM, pH 8) with 5% DMSO as a co-solvent to an initial concentration of 5 µM

and placed in borosilicate test tubes. In all experiments, an ionic strength of 0.5 M

was maintained by complmenting other ions (e.g. NaCl) with NaClO4 when

necessary. For analysis, entire samples were sacrificed at each time point. Samples

were then placed in centrifuge tubes and pyrene-d10 was added to the solution as a

surrogate. Next, an extraction was performed using hexane (20 mL) and samples were

Page 112: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  107  

sonicated for 20 min. The organic layer was separated by centrifugation (3000 rpm, 3

min) and passed through filter paper to remove residual water. The hexane sample

was removed by rotary evaporation and blown to dryness with nitrogen. Lastly, the

samples were re-dissolved in 100 µL of hexane, which was pre-spiked with

phenanthrene-d10 and analyzed by GC/MS.

The system used for transient absorption (laser flash photolysis) experiments

has been previously described in detail21. Briefly, the setup used was an EOS transient

absorption spectroscopy system (Ultrafast Systems, Sarasota, USA). Excitation pulses

were generated by converting the 795 nm output of a Solstice regeneratively

amplified Ti:sapphire laser (Spectra-Physics, Darmstat, Germany) to the desired

wavelength using a TOPAS-C (Light Conversion, Lithuania) optical parametric

amplifier. The excitation wavelength used was 333 nm, a local absorption maximum

for PYR. Laser powers of 4-8 mW were used. At higher laser powers, the production

of PYR!+ through two-photon absorption is more pronounced, yielding stronger

signals. For each experiment, 200 mL of a 30 µM PYR solution (70% methanol 30%

H2O) was placed in a round-bottom flask and purged with synthetic air. Using a

peristaltic pump, the solutions were continuously cycled through a 10-mm quartz

flow-through cuvette. To these solutions, the various electron donors and nucleophiles

were spiked in at different concentrations to determine their reactivity.

Results and Discussion

In air-equilibrated aqueous solutions containing 0.5-0.2M NaCl, PYR

degraded slowly, with a half-life of about 3.5 hrs when irradiated with light bulbs

emitting wavelengths from 340 to 400 nm. Only the tail of the PYR UV absorbance

overlaps with this light source, and under such conditions, two-photon absorption is

unlikely. In all cases, the formation of Cl-PYR was observed, with apparent yields of

Page 113: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 108  

around 5%. These results are consistent with earlier work20 showing PYR and other

PAHs can become chlorinated when irradiated in artificial seawater. Our proposed

reaction mechanism is depicted in Figure 2. We hypothesize that Cl-PAH formation is

initiated by the one-electron oxidation of photoexcited singlet-state PYR by O2, which

yields PYR!+ and superoxide O2!-. In the next step, Cl- acts as a weak nucleophile and

attacks the delocalized carbocation at position 1 in PYR!+. Nucleophilic attack at

position 1 is known to occur2, and can be rationalized by the fact that this location has

the highest loading of positive charge to minimize positive-positive interactions.

Following nucleophilic attack, another one-electron oxidation by O2 followed by loss

of a proton yields the end product, Cl-PYR.

Page 114: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  109  

Figure 3- The degradation of pyrene (top) and the resulting formation of 1-chloropyrene (bottom) under various conditions.

As noted above, this mechanism relies on the presence of O2 to act as an

oxidant to both form PYR!+ and act as the terminal acceptor of the second electron in

a subsequent step. Therefore, removal of O2 would be expected to eliminate Cl-PYR

formation as well as slow the overall degradation of PYR. To test for the importance

of O2, a series of samples were purged with Ar and sealed during photolysis. Removal

of oxygen slows the degradation of PYR considerably and the generation of Cl-PYR

is no longer observed (Figure 1). These results also agree with previous work that has

0 100 200 3000

5

10

15

20

25

30

35 Cl 0.8 M Cl 0.5 M Cl 0.5 M DABCO Cl 0.5 M Ar purged

Cl-P

yren

e (n

M)

Time (min)

0 100 200 3001.5

2.0

2.5

3.0

3.5

4.0

Pyr

ene

(µM

)

Time (min)

Cl 0.8 M Cl 0.5M Cl 0.5 M DABCO Cl 0.5 M Ar purged

Page 115: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 110  

shown select PAHs, including PYR, appear to degrade primarily via a radical cation

(PYR!+) formed from a one-electron oxidation by O2 following photoexcitation22-24.

This oxidation appears to take place from the singlet state of PYR through collisional

charge-transfer complexes with O2. Oxidation from the triplet state is not likely, as

demonstrated by others22,23 and in agreement with our observations by transient

absorbance measurements showing the growth of the PYR!+ signal on a timescale

independent of that of triplet PYR (figure 3). The absence of Cl-PYR formation in

these experiments is also significant as it is inconsistent with photoexcited PYR as

being capable of sensitizing the production of any reactive form of chloride, without

an additional oxidant (O2) present.

Figure 4- The proposed reaction scheme leading to the photochemical formation of 1-

chloropyrene.

To further test for the involvement of PYR!+ as a key intermediate in Cl-PYR

formation, we performed a series of photolysis experiments with an electron donor

present to see if a decrease in Cl-PYR formation would be observed. PYR solutions

containing DABCO (100 µM) showed a large reduction in Cl-PYR production. Our

initial expectation was that the presence of DABCO, a good electron donor, would

lead to enhanced regeneration of PYR from PYR!+ and, in addition to reducing Cl-

PYR formation, would slow the observed rate of PYR degradation. In Figure 1 we see

that PYR degradation is only slowed to a minor degree compared to the loss of Cl-

PYR formation. This is likely the result of DABCO acting not only as an electron

donor, but also as a nucleophile capable of reacting with PYR!+.

Pyr h! 1Pyr* PyrO2 O2

+ Cl- Cl-PyrO2 O2

-H+Cl-Pyr

Page 116: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  111  

In order to test the nucleophilic behavior of Cl- in this reaction, we performed

a set of competition experiments with other nucleophiles. When two nucleophiles are

present in solution at the same concentration together and able to compete for reaction

with a cation, their respective product ratios can be reasonably predicted based on the

relative strength of each nucleophile assuming they react in a similar manor. For this

test, photolysis experiments were performed with PYR solutions containing equal

amounts of Cl- and either Br- or CN-, both known to be stronger nucleophiles than Cl-

25. In the presence of Br- a large reduction of Cl-PYR formation was seen and 1-

bromopyrene (Br-PYR) was formed in relatively large quantities. The product ratio,

as determined by the initial rates of formation for Cl-PYR and Br-PYR was

approximately 10. This ratio agrees quite well with observations of the relative

strengths of these nucleophiles in other aqueous systems26.

Transient absorption experiments were also performed, allowing direct

observation of PYR!+ and other PYR transient species. Laser flash photolysis has

been used extensively in the past to observe22,27 the behavior and reactivity of PAHs

and their various radical species, and can be a suitable technique to determine reaction

rate constants. Our main objective was to obtain the second-order reaction rate

constant for the addition of Cl- to PYR!+, however this reaction occurs too slowly to

affect the decay of the PYR!+ signal, even at Cl- concentrations of 1M. Attempts to

perform these experiments at Cl- concentrations above 1M resulted on poor transient

signals of PYR!+, most likely due to solubility issues. Experiments performed with Br-

also showed no noticeable change in the decay lifetime of PYR!+. Based on the lack

of a distinguishable change in lifetime at a Br- concentration of 1M, we can put an

upper limit on the bimolecular reaction rate constant at 5x10-5 M-1 s-1. This is in good

agreement with other studies of the rates of reactions of Br- as a nucleophile. We also

attempted to observe quenching of PYR!+ by electron donors. In these experiments we

Page 117: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 112  

observed (Figures 3 and 4) that N3- reacted with PYR!+ at or near diffusion controlled

rates. Despite not being able to observe the reaction between Cl- and PYR!+, all of

these findings are in good agreement with the observations of others and support our

proposed reaction mechanism.

Figure 3 – 3D Transient absorption spectra of PYR. In the left panel the sample was purged with O2 and contained only PYR. The absorption of PYR!+ can be seen at 455 nm. In the right panel, the sample was purged with air and contained 100mM N3

-, and the PYR!+ is absent.

Page 118: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  113  

Figure 4- Transient absorption decay traces of PYR!+ (ΔAbs @ 455nm) in air purged samples collected in the absence (blue crosses) and presence (red x) of the electron donor N3

-. The fast component seen in both traces is the decay of 1PYR* which also absorbs at 455nm.

As mentioned before, the halogenation of PAHs, including PYR can occur

through a variety of mechanisms, which largely depend on which oxidants are present

in the system. During combustion, high temperatures give rise to gas-phase radical

reactions, which are well known to form halogenated organic compounds when an

appropriate halogen source is available. In solution, PAHs have been shown to be

chlorinated by Cl2 and ClO- in a non-photochemical manor and thus through a

different mechanism than the one proposed here. Under well-lit conditions, and with

Cl- as the most freely available chlorine source, the formation of Cl-PYR in marine

systems is unlikely to involve any other reactive form of Cl, such as ClO- or Cl2!-. We

argue based on our data and earlier findings13,20 that in marine environments a

significant contribution of Cl-PAHs present are of photochemical origin.

0 20 40 60 80 1000.0

5.0m

10.0m

15.0m

20.0m

25.0m

30.0m

Δ A

bsor

banc

e

Time (µs)

Page 119: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 114  

References (1) Cavalieri, E.; Rogan, E. Role of Radical Cations in Aromatic Hydrocarbon Carcinogenesis. Environmental Health Perspectives 1985, 64, 69-84. (2) Bostrom, C. E.; Gerde, P.; Hanberg, A.; Jernstrom, B.; Johansson, C.; Kyrklund, T.; Rannug, A.; Tornqvist, M.; Victorin, K.; Westerholm, R. Cancer risk assessment, indicators, and guidelines for polycyclic aromatic hydrocarbons in the ambient air. Environmental Health Perspectives 2002, 110, 451-488. (3) Umbuzeiro, G. A.; Franco, A.; Martins, M. H.; Kummrow, F.; Carvalho, L.; Schmeiser, H. H.; Leykauf, J.; Stiborova, M.; Claxton, L. D. Mutagenicity and DNA adduct formation of PAH, nitro-PAH, and oxy-PAH fractions of atmospheric particulate matter from Sao Paulo, Brazil. Mutat. Res. Genet. Toxicol. Environ. Mutagen. 2008, 652, 72-80. (4) Diamond, M. L.; Gingrich, S. E.; Fertuck, K.; McCarry, B. E.; Stern, G. A.; Billeck, B.; Grift, B.; Brooker, D.; Yager, T. D. Evidence for Organic Film on an Impervious Urban Surface:   Characterization and Potential Teratogenic Effects. Environmental Science & Technology 2000, 34, 2900-2908. (5) Wills, L. P.; Zhu, S.; Willett, K. L.; Di Giulio, R. T. Effect of CYP1A inhibition on the biotransformation of benzo[a]pyrene in two populations of Fundulus heteroclitus with different exposure histories. Aquatic Toxicology 2009, 92, 195-201. (6) Mahmoudi, N.; Porter, T. M.; Zimmerman, A. R.; Fulthorpe, R. R.; Kasozi, G. N.; Silliman, B. R.; Slater, G. F. Rapid Degradation of Deepwater Horizon Spilled Oil by Indigenous Microbial Communities in Louisiana Saltmarsh Sediments. Environmental Science & Technology 2013, 47, 13303-13312. (7) Boule, P.; Pagni, R.; Sigman, M. In Environmental Photochemistry; Springer Berlin Heidelberg: 1999; Vol. 2 / 2L, p 139-179. (8) Haglund, P.; Alsberg, T.; Bergman, A.; Jansson, B. Analysis of halogenated polycyclic aromatic hydrocarbons in urban air, snow and automobile exhaust. Chemosphere 1987, 16, 2441-2450. (9) Horii, Y.; Ok, G.; Ohura, T.; Kannan, K. Occurrence and Profiles of Chlorinated and Brominated Polycyclic Aromatic Hydrocarbons in Waste Incinerators. Environmental Science & Technology 2008, 42, 1904-1909. (10) Ohura, T.; Horii, Y.; Kojima, M.; Kamiya, Y. Diurnal variability of chlorinated polycyclic aromatic hydrocarbons in urban air, Japan. Atmospheric Environment 2013, 81, 84-91. (11) Ni, H.-G.; Guo, J.-Y. Parent and Halogenated Polycyclic Aromatic Hydrocarbons in Seafood from South China and Implications for Human Exposure. Journal of Agricultural and Food Chemistry 2013, 61, 2013-2018. (12) Horii, Y.; Ohura, T.; Yamashita, N.; Kannan, K. Chlorinated Polycyclic Aromatic Hydrocarbons in Sediments from Industrial Areas in Japan and the United States. Archives of Environmental Contamination and Toxicology 2009, 57, 651-660. (13) Sankoda, K.; Kuribayashi, T.; Nomiyama, K.; Shinohara, R. Occurrence and Source of Chlorinated Polycyclic Aromatic Hydrocarbons (Cl-PAHs) in Tidal Flats of the Ariake Bay, Japan. Environmental Science & Technology 2013, 47, 7037-7044. (14) Horii, Y.; Khim, J. S.; Higley, E. B.; Giesy, J. P.; Ohura, T.; Kannan, K. Relative Potencies of Individual Chlorinated and Brominated Polycyclic Aromatic Hydrocarbons for Induction of Aryl Hydrocarbon Receptor-Mediated Responses. Environmental Science & Technology 2009, 43, 2159-2165.

Page 120: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  115  

(15) Mori, Y.; Goto, S.; Onodera, S.; Naito, S.; Matsushita, H. Aqueous chlorination of tetracyclic aromatic hydrocarbons: reactivity and product distribution. Chemosphere 1991, 22, 495-501. (16) Hu, J.; Jin, X.; Kunikane, S.; Terao, Y.; Aizawa, T. Transformation of Pyrene in Aqueous Chlorination in the Presence and Absence of Bromide Ion:   Kinetics, Products, and Their Aryl Hydrocarbon Receptor-Mediated Activities. Environmental Science & Technology 2005, 40, 487-493. (17) Shiraishi, H.; Pilkington, N. H.; Otsuki, A.; Fuwa, K. Occurrence of chlorinated polynuclear aromatic hydrocarbons in tap water. Environmental Science & Technology 1985, 19, 585-590. (18) Sugiyama, H.; Katagiri, Y.; Kaneko, M.; Watanabe, T.; Hirayama, T. Chlorination of pyrene in soil components with sodium chloride under xenon irradiation. Chemosphere 1999, 38, 1937-1945. (19) Sankoda, K.; Nomiyama, K.; Kuribayashi, T.; Shinohara, R. Halogenation of Polycyclic Aromatic Hydrocarbons by Photochemical Reaction under Simulated Tidal Flat Conditions. Polycyclic Aromatic Compounds 2013, 33, 236-253. (20) Sankoda, K.; Nomiyama, K.; Yonehara, T.; Kuribayashi, T.; Shinohara, R. Evidence for in situ production of chlorinated polycyclic aromatic hydrocarbons on tidal flats: Environmental monitoring and laboratory scale experiment. Chemosphere 2012, 88, 542-547. (21) Wenk, J.; Eustis, S. N.; McNeill, K.; Canonica, S. Quenching of Excited Triplet States by Dissolved Natural Organic Matter. Environmental Science & Technology 2013, 47, 12802-12810. (22) Fasnacht, M. P.; Blough, N. V. Mechanisms of the Aqueous Photodegradation of Polycyclic Aromatic Hydrocarbons. Environmental Science & Technology 2003, 37, 5767-5772. (23) Sigman, M. E.; Schuler, P. F.; Ghosh, M. M.; Dabestani, R. T. Mechanism of Pyrene Photochemical Oxidation in Aqueous and Surfactant Solutions. Environmental Science & Technology 1998, 32, 3980-3985. (24) Fasnacht, M.; Blough, N. Kinetic analysis of the photodegradation of polycyclic aromatic hydrocarbons in aqueous solution. Aquatic Sciences 2003, 65, 352-358. (25) Schwarzenbach, R. P.; Gschwend, P. M.; Imboden, D. M. Environmental Organic Chemistry; 2nd ed.; John Wiley & Sons: Hoboken, NJ, 2003. (26) Hine, J. Physical Organic Chemistry; McGraw-Hill: New York, 1962. (27) Steenken, S.; Warren, C. J.; Gilbert, B. C. Generation of radical-cations from naphthalene and some derivatives, both by photoionization and reaction with SO4-[radical dot]: formation and reactions studied by laser flash photolysis. Journal of the Chemical Society, Perkin Transactions 2 1990, 335-342.

Page 121: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 116  

Page 122: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  117  

Chapter 5

Conclusions and outlook

Page 123: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 118  

The work presented in this thesis addressed several topics in the context of the

environmental photochemistry of organic pollutants. In Chapter 2, the direct

photochemical behavior of three selected OH-PBDEs was studied. We found that

each of the studied congeners transformed photochemically to yield new products that

are themselves, known environmental pollutants. The formation of toxic PBDDs from

OH-PBDEs was observed in all cases, and illustrates the need to further investigate

the fate of organic micropollutants. We showed that the application of fundamental

chemical knowledge can explain the results and help predict the environmental

behavior of these organic micropollutants. For each of the individual congeners, we

identified structural differences, e.g., differences in their bromination patterns, which

accounted for the differences in their transformation fates. For instance, one of the

transformation pathways common to each congener was the cleavage of the ether

linkage to yield brominated phenols. The production of bromophenol from each

congener correlated with the pKa value of the resulting bromophenol, indicating the

efficiency of this transformation pathway is likely a function of the bromophenol to

act as a leaving group. This case highlights that, by understanding the reaction

mechanism, estimates can be made about the environmental fate of other compounds

using simple chemical parameters. PBDD formation yields from each OH-PBDE

congener were found to differ, and the congener that formed PBDD with the highest

efficiency was the natural product, 6-OH-PBDE 99. This result is interesting in light

of the fact that the majority PBDDs found in Baltic Sea sediments and biota samples

appear to be primarily natural in origin. Our work lends evidence to this conclusion

by demonstrating a plausible photochemical route from a known natural product, to a

PBDD found in environmental samples.

Page 124: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  119  

In Chapter 3, the oxidation of phenols and anilines by triplet state MB was

investigated. In the case of anilines, this reaction was observed to take place via a

one-electron transfer reaction to yield an aniline radical cation and neutral radical MB

species. The reaction with phenols, however proceeded by a proton-coupled coupled

electron transfer to form a neutral phenoxyl radical and a protonated MB radical

cation. These results were supported both by the characteristic spectroscopic

signatures of each MB reaction product observed by transient absorption studies, but

also by other experiments, including ones that probed the kinetic isotope effect of the

O-H bond strength in phenols. Empirical relationships between the reaction free

energy and the observed rate constants were established for both types of reaction

mechanisms. In the case of the anilines, the reaction rate constants could be modelled

using the reaction free energy calculated from the aniline oxidation potential, and

were fit using a Sandros-Boltzmann type dependence on free energy. Reaction rate

constants of phenols correlated with the phenolic O–H BDFE, and could be modelled

by a multi-parameter equation requiring only the BDFE and pKa values of the

phenols. By combining reaction rate constants determined from transient absorption

measurements with computationally derived oxidation potentials, this work explored a

framework for making better predictions about the indirect photochemical role in the

fate of some pollutant classes.

In Chapter 4, an investigation was performed on the mechanism of

photochemical Cl-PYR formation from PYR in aqueous solution. Previous studies

have shown a correlation between concentrations of Cl-PYR in marine sediments and

water salinity. In the course of this study, we collected data to test our hypothesis that

Cl-PYR formation was a result of the nucleophilic attack by Cl- on PYR radical

cations formed after the oxidation of photoexcited PYR by O2. In the presence of an

added electron donor N3-, Cl-PYR formation by photolysis was suppressed. This

Page 125: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 120  

observation is likely the result of N3- quenching the radical cations before chloride is

able to attack. This idea is supported by transient absorption measurements, in which

the PYR radical cation signal was eliminated by the addition of N3- and S2O3

2-. The

role of O2 in oxidizing photoexcited PYR was examined by purging samples with Ar

prior to photolysis. Results of these experiments showed Cl-PYR was not formed in

the absence of O2 and the observed rate of pyrene degradation was decreased. Finally,

the nucleophilic behavior of Cl- was examined through competition experiments in

which PYR was photolyzed in the presence of Cl- and an additional nucleophile. With

Br- added as the competing nucleophile, the relative amounts of Br-PYR and Cl-PYR

formed matched what was expected for a nucleophilic reaction mechanism. This work

provides additional evidence to support the hypothesis that some Cl-PAHs found in

contaminated marine environments are formed in situ.

Outlook

For each topic investigated in this thesis, there exist a number of additional

questions and problems that are worthy of future research. An interesting question

raised by our results showing that OH-PBDEs may be responsible for the presence of

PBDDs in the Baltic Sea is whether or not this photochemical process can happen

within marine biota that produce OH-PBDEs. Certain marine species, including

sponges and red algae that are found in the Baltic Sea are known to produce OH-

PBDEs. In addition to containing OH-PBDEs, samples of some these species have

also been shown to contain high concentrations of PBDDs. Whether or not PBDDs

are produced from a biosynthetic pathway or not us unclear, and the connection

between OH-PBDEs and PBDDs in marine biota is an open question. The possibility

exists then that the PBDDs found in marine creatures exposed to adequate sunlight

could be generated photochemically from OH-PBDEs. To address this question, more

Page 126: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  121  

detailed analytical studies could be performed to look for matches in the expected and

identified congener profiles of OH-PBDEs and PBDDs in these samples. Another

possible approach would be controlled light-exposure experiments, to access if

exposure to the wavelengths that lead to PBDD formation has an effect on their

concentration. This area of research is unique, because it reminds us that in some

cases what seems like a pollutant may actually be an interesting part of the chemistry

of life.

The ability to predict the environmental fate and effects of organic compounds

is perhaps the single biggest goal of environmental chemistry. To meet this goal more

sophisticated and reliable prediction methods will be required. Our attempt at

modeling aniline and phenol degradation rates with MB represents an important step

in developing more accurate predictions of the role of 3CDOM in pollutant

degradation. The use of MB as our model triplet oxidant, while justified in the context

of Chapter 3, makes comparing our results to the expected behavior of 3CDOM as an

oxidant difficult. Future studies should focus on more difficult to oxidize anilines and

the use of excited state acceptors. One set of acceptors that may be important that

have received little attention are quinones. Like the more well-studied aromatic

ketones, some quinones exhibit high triplet yields and may also play a role in the

photoinduced oxidation of pollutants. Because 3CDOM contains a mix of

photoexcited molecules able to participate in oxidation reactions, data from a wide

range of excited acceptors will need to be incorporated into any prediction model.

Lastly, it would be interesting to see of other examples of the

photohalogenation mechanism described in Chapter 4 can be found in marine

environments. A number of compounds are known to degrade via a one-electron

oxidation or photoionization mechanism, which leaves the possibility open for other

pollutants to be chlorinated in the same manor as PYR in seawater. Another area that

Page 127: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

 122  

could be addressed is the relative importance of in situ chlorination to other Cl-PAH

sources. By analyzing substitution patterns of selected Cl-PAHs, which can indicate

radical vs. cationic formation, it may be possible to determine the overall

contributions from different sources.

Page 128: Research Collection Doctoral Thesis Transformation mechanisms of organic micropollutants via

  123  

Acknowledgements I would like to thank my advisor, Kris, who I first had the opportunity to meet when I worked for him as an undergraduate student the summer before my senior year. His enthusiasm for science and interest in his students is what convinced me to return to Minnesota, and subsequently move with the lab to ETH. Thank you Kris, for the opportunity to be a part of an engaging and welcoming research team and helping to me to learn what it means to “push back the frontiers of science”. The fellow students are what made my time in the McNeill lab truly memorable. As a young grad student, I remember thinking how nice it was to work in such a positive atmosphere. As the years passed and I became one of the veteran students, I did my best to be the same example of a good lab mate as the students that came before me were. From my first group happy hour on the patio at Burrito Loco in Dinkytown to the shore of the Seealpsee after an excellent group hike in Appenzell, I am thankful for the good times spent in the company of great people. Sarah K, I am grateful that you and your knowledge were around for so long while I was in this group. Our conversations about the projects we were working on were almost as good as the many others we had about life outside of the lab. Soren, Thank you for taking me under your wing and introducing me to the wonderful world of time-resolved spectroscopy. You left big shoes to fill (literally!) as the overseer of the McNeill laser lab, and I enjoyed the months we spent working together. Rebekka and Lilli, thank you for translating my dissertation summary. After four years in Switzerland you would think I could handle this, but with words such as “Elektronenübertragungsechanismus“ it was nice to have the help. Thank you Mom and Dad, for fostering my interest in science and encouraging me to always be curious. Every step of the way, from buying me my much-loved microscope to helping me with a science fair project using a real dynamometer, you were always there to enable my scientific mind. You were, and always will be wonderful parents. Jessica, I cannot thank you enough for all that you do. I am so grateful to have had you along for this adventure that took us from Florida to Minnesota and on to Zurich. No matter what direction we travelled, it always brought me closer to you. Having a loving person and now family to come home to every day makes my life complete. I love you so much and I look forward to whatever new adventures the future brings our way.


Recommended