39
New pyrrole-based ligand for first row transition metal complexes Degree Project C in Chemistry (1KB010) thesis by: Ricardo J. Fernández Terán Supervisor: Dr. Anders Thapper Subject Specialist: Dr. Lukasz Pilarski Examiner: Prof. Dr. Helena Grennberg May 2015

New pyrrole-based ligand for first row transition metal complexes

  • Upload
    vutruc

  • View
    236

  • Download
    5

Embed Size (px)

Citation preview

New pyrrole-based ligand for first row

transition metal complexes

Degree Project C in Chemistry (1KB010) thesis by:

Ricardo J. Fernández Terán

Supervisor: Dr. Anders Thapper

Subject Specialist: Dr. Lukasz Pilarski

Examiner: Prof. Dr. Helena Grennberg

May 2015

2

To the loving memory of Prof. Dr. Isaura De Sanctis,

who passed away in December 2014,

for all what I learned from her,

including the love and passion for

Inorganic Chemistry.

3

LIST OF ABBREVIATIONS

acac Acetylacetonate (Pentane-2,4-dionate) ligand

Boc tert-butyloxycarbonyl (protecting group)

bpy 2,2’-bipyridine

DEPT Distortionless Enhancement by Polarisation Transfer (NMR pulse sequence)

DMAP 4-(dimethylamino)pyridine

Et2O Diethyl ether

Et3N Triethylamine

EtOH Ethanol

HTMP 2,2,6,6-tetramethylpiperidine

LiTMP Lithium 2,2,6,6-tetramethylpiperidide

MeOH Methanol

NADP+ Nicotinamide adenine dinucleotide phosphate

NADPH Nicotinamide adenine dinucleotide phosphate (protonated)

n-BuLi n-butyllithium

PCET Proton coupled electron transfer

PY Pyridine

PY5 2,6-bis(methoxydi(pyridin-2-yl)methyl)pyridine

PY5-OH Pyridine-2,6-diylbis(di(pyridin-2-yl)methanol)

PyPY2 5-(methoxydi(pyridin-2-yl)methyl)-pyrrole-2-carboxylate

PyPY4 (1H-pyrrole-2,5-diyl)bis(di(pyridin-2-yl)methanol)

RC Radical coupling

t-BuLi tert-butyllithium

THF Tetrahydrofuran

TON Turnover number

TOF Turnover frequency

WNA Water nucleophilic attack

WOC Water oxidation catalyst

4

ABSTRACT

The negative impact of the use of traditional fossil fuels in the environment and their limited

nature has prompted research in new, alternative sources of energy. Artificial photosynthesis,

which comprises the splitting of water into hydrogen and oxygen, can be used as clean source of

energy in many applications, especially since the main energy source is the sun. The oxidation

of water into dioxygen by transition metal complexes invokes the necessity for new catalysts to

be synthesized and studied, looking forward to achieving catalytic activities closer to those

shown by the oxygen evolving complex in Photosystem II. In this project, a new tridentate

pyrrole-based ligand and its metal complexes with Copper(II) and Nickel(II) were synthesized.

The obtained ligand, methyl 5-(methoxydi(pyridin-2-yl)methyl)-1H-pyrrole-2-carboxylate, was

synthesized in moderate yields (38%). Differences in reactivity between pyridine and pyrrole

diesters were also explored, since the designed pentadentate ligand, (1H-pyrrole-2,5-

diyl)bis(di(pyridin-2-yl)methanol)), could not be obtained. Also, structural differences where

observed in the products when following similar synthetic schemes for both ester substrates.

RESUMEN

El impacto negativo sobre el ambiente que tiene el uso de combustibles fósiles y su naturaleza

limitada han promovido la investigación en fuentes nuevas y alternativas de energía. La

fotosíntesis artificial, que involucra la ruptura de la molécula de agua en oxígeno e hidrógeno,

puede ser usada como una fuente limpia de energía para muchas aplicaciones, especialmente

debido a que la principal fuente de energía es el sol. La oxidación de agua en oxígeno molecular

mediada por complejos de metales de transición invoca la necesidad de sintetizar y estudiar

nuevos catalizadores, teniendo como meta la obtención de actividades catalíticas que se

acerquen a las exhibidas por el complejo liberador de oxígeno en el Fotosistema II. En este

proyecto se sintetizó un nuevo ligando tridentado basado en pirrol y sus complejos con Cobre(II)

y Níquel(II).

El ligando obtenido, metil 5-(metoxidi(2-piridinil)metil)-1H-pirrol-2-carboxilato, fue sintetizado

con un rendimiento moderado (38%) mediante el uso de reactivos de organolitio. Las diferencias

en la reactividad de los diésteres de piridina y pirrol fueron exploradas, debido a que el ligando

pentadentado diseñado, (1H-pirrol-2,5-diil)bis(di(2-piridil)metanol), no pudo ser obtenido. De

igual forma, se pudieron observar algunas diferencias estructurales en los productos cuando se

siguieron esquemas sintéticos similares para ambos sustratos.

5

TABLE OF CONTENTS

New pyrrole-based ligand for first row transition metal complexes ............................................ 1

LIST OF ABBREVIATIONS .................................................................................................. 3

ABSTRACT ............................................................................................................................... 4

RESUMEN ................................................................................................................................ 4

TABLE OF CONTENTS ......................................................................................................... 5

Introduction ..................................................................................................................... 7

The global energy problem .......................................................................................... 7

Natural and artificial photosynthesis ......................................................................... 7

Nature’s choice: Mn4Ca redox core in Photosystem II ................................................ 8

First attempts at metal-catalysed water oxidation..................................................... 9

Water oxidation with mononuclear Iron and Cobalt complexes ............................... 10

Previous catalysts based in pentadentate polypyridine ligands............................... 12

Synthetic routes for polypyridine ligands: considerations ........................................ 13

Proposed modifications to the ligand scaffold ........................................................... 14

Synthetic approach to the proposed ligand ............................................................... 15

Aim of the project .......................................................................................................... 16

Experimental .................................................................................................................. 16

Chemicals and solvents ............................................................................................. 16

Spectroscopic characterisation .................................................................................. 16

Synthesis of pyrrole-2,5-dicarboxylic acid diesters (dimethyl, diethyl) .................... 16

Synthesis of Di(n-butyl) pyrrole-2,5-dicarboxylate ................................................... 17

Synthesis of N-Boc-pyrrole (tert-butyl pyrrole-1-carboxylate) .................................. 17

Synthesis of 1-(tert-butyl) 2,5-dimethyl pyrrole-1,2,5-tricarboxylate ....................... 17

Synthesis of methyl 5-(methoxydi(pyridin-2-yl)methyl)-pyrrole-2-carboxylate

(PyPY2) ................................................................................................................................. 18

Synthesis of Nickel(II) and Copper(II) complexes with PyPY2 ................................ 19

Results and Discussion ................................................................................................ 19

Synthesis of pyrrole-2,5-dicarboxylic acid diesters (dimethyl, diethyl) .................... 19

6

Synthesis of Di(n-butyl) pyrrole-2,5-dicarboxylate ................................................... 20

Synthesis of N-Boc-pyrrole ........................................................................................ 21

Synthesis of 1-(tert-butyl) 2,5-dimethyl pyrrole-1,2,5-tricarboxylate ....................... 21

Synthesis of methyl 5-(methoxydi(pyridin-2-yl)methyl)-pyrrole-2-carboxylate

(PyPY2) ................................................................................................................................. 22

Attempted synthesis of Nickel(II) and Copper(II) complexes of the PyPY2 ligand . 23

Conclusions .................................................................................................................... 25

Acknowledgements ....................................................................................................... 26

References ...................................................................................................................... 27

Appendixes ..................................................................................................................... 31

1H NMR spectrum of N-Boc pyrrole .......................................................................... 32

1H, 13C and DEPT135 NMR spectra of 1-(tert-butyl) 2,5-dimethyl pyrrole-1,2,5-

tricarboxylate ....................................................................................................................... 33

1H, 13C, DEPT135 and DEPT90 NMR spectra of PyPY2 .......................................... 35

1H NMR spectrum of methyl 1H-pyrrole-2-carboxylate ........................................... 39

7

INTRODUCTION

The global energy problem

The current global energy problem, the need for alternative sources of energy and the fact that

traditional use of fossil fuels and other non-renewable energy sources have a negative impact

on the climate and the environment1-4, has promoted the development of new renewable and

cleaner energy sources. Among them, one of the most theoretically promising is the conversion

of solar energy into chemical energy through the light-driven splitting of water, also known as

artificial photosynthesis, which generates oxygen gas and hydrogen, a carbon-free fuel with the

highest energy output relative to molecular weight.

Artificial photosynthesis thus presents an environmentally friendly alternative to the use of

fossil fuels, and the development of new molecular catalysts for artificial photosynthesis is an

active field of research with possible many applications.

Natural and artificial photosynthesis

The natural photosynthetic process consists of two coupled processes: water oxidation and CO2

fixation via proton reduction, to form NADPH, which is then used to synthesize sugars and more

complex biomolecules.

The light-dependent reactions of photosynthesis occur mainly on two protein complexes:

Photosystem II (containing the oxygen-evolving complex) and Photosystem I (responsible for the

reduction of NADP+ to NADPH). This process is shown in Fig. 1:

+1.2

+0.8

+0.4

0.0

-0.4

-0.8

-1.2

2H O2

Mn Ca4

TyrZ

O + 4H2+ 4e−

ChlP680

ExcitedChl P680*

Photo

syste

m II Pheo

QAQB

PQFeS

Cyt.fPC

Cyt. b & b6H 6L

2H+

2H+ ChlP700

ExcitedChl P700*

Photo

syste

m I

A0A1

FX

F /FA B

FDFNR

NADPH

NADP+

LightEnergy

LightEnergy

Fig. 1: Light-dependent reactions of photosynthesis (Z-scheme): roles of Photosystems I and II in the different

processes. Orange arrows illustrate the electron transfer chains, formed by various cofactors.

Adapted from ref. 5 with permission.

8

The oxidation of water is catalysed in Photosystem II by a redox-active structure that contains

a tetramanganese-calcium cluster (Mn4Ca), bound with a Tyrosine residue (named TyrZ); this

oxygen-evolving complex binds two water molecules and stores the four oxidizing equivalents

that are required to drive the oxidation of water. In the light-independent reactions, CO2 is

reduced by NADPH to ultimately yield sucrose, starch and cellulose after a series of reactions

(known as the Calvin-Benson cycle). These sugars are the main compounds responsible for

natural solar energy storage5. Both water oxidation and proton reduction (H+ to H2(g)) can be

achieved by means of artificial catalysts (for a very complete and recent review about WOCs see

ref. 6). These reactions can be carried out both by using a chemical redox agent or

photochemically, by using a photosensitizer, which acts as a redox mediator by photon induced

electron transfer processes.

In a process mimicking natural photosynthesis, the use of solar energy to extract electrons and

release protons from water is called photosynthetic water oxidation or oxygen evolution, and

development of new catalysts for this half reaction is of great interest for the field, since water

oxidation is a complicated reaction involving proton-coupled electron transfer (PCET) of four

electrons, yielding four protons and two oxygen molecules from two water molecules (Eq. 1).

2H2O ⟶ O2 + 4H+ + 4e− (𝐸0 = −1.23 𝑉 vs. SHE) Eq. 1

This reaction is thermodynamically uphill and pH dependant (since protons are being

produced), and to be carried out separately from proton reduction or other catalytic reduction

systems, a sacrificial electron acceptor is required: typically peroxydisulfate anion (S2O82−) is

used for this purpose.

Nature’s choice: Mn4Ca redox core in Photosystem II

The natural catalyst, a Mn4Ca reactive centre inside Photosystem II is shown in Fig. 2:

Fig. 2: Structure of the Mn4Ca reactive centre in Photosystem II. Adapted from ref. 7 with permission.

9

The structure of this reactive centre was determined by several X-ray diffraction and absorption

spectroscopic techniques7-11. This Mn4Ca core has been the base for many artificial catalysts,

aimed to imitate, model and understand the structure-activity relationship and the role of the

different components in the high catalytic activity of the natural catalyst, with the ultimate goal

of reproducing this catalytic activity in a synthetic complex that can be later inserted into a

water-splitting device.

Very recently, a synthetic catalyst mimicking the oxygen-evolving centre of photosynthesis was

reported by the group of Zhang et al.12, 13. This synthetic analogue consists of a Mn4Ca core with

ligands derived from carboxylic acids and has a very close structural resemblance to the natural

one. However, its catalytic activity is much lower than the natural catalyst, which is a matter

that is still under investigation. The structure of this artificial catalyst is shown in Fig. 3:

Fig. 3: Structures of the natural and artificial (synthetic) catalysts. See caption in figure.

Reprinted from ref. 13 with permission.

The natural turnover number (TON) and turnover frequency (TOF) of the PSII catalyst have

been estimated to be around 600.000 mol O2 (mol cat)−1 and 40 s−1, respectively14, 15.

First attempts at metal-catalysed water oxidation

The first successful attempts at metal-catalysed water oxidation date back to a publication from

1982, on which Gersten et al.16 reported the catalytic oxidation of water by an oxo-bridged

ruthenium dimer, cis,cis-[(bpy)2(OH2)RuIII-O-RuIII(OH2)(bpy)2]4+, later known in the field as the

“blue dimer”, which was the only reported molecular catalyst for water oxidation for more than

a decade.

10

Many other catalysts based on polynuclear manganese and ruthenium systems have been

reported, and it was in 2012 when Duan et al. reported a pair of molecular ruthenium catalysts

with water-oxidation activities comparable to that of the natural Photosystem II17. Several other

complexes have been reported in the last decade, with the notable examples of new cobalt and

iridium-based systems18. The structures of these catalysts are summarized in Fig. 4, which

illustrates the chronological evolution of representative water oxidation catalysts among the

past decades.

Fig. 4: Chronological listing of representative WOCs reported over the past three decades.

Reprinted from ref. 18 with permission.

Water oxidation with mononuclear Iron and Cobalt complexes

Despite the progress made in the synthesis of highly active polynuclear metal complexes for

water oxidation catalysis, the fact that ruthenium is a fairly expensive and non-abundant metal

limits the future application of these catalysts as a possible solution to the global energy

problem. The use of earth-abundant metals like iron, cobalt and manganese for water oxidation

catalysis has been an active field of investigation, on which fewer catalytic systems have been

reported so far. Apart from the use of earth-abundant metals, there are two other important

aspects to consider: the first downside aspect of polynuclear metal complexes that makes them

unattractive for a big scale application is atom economy: several metallic sites involve a higher

cost in production and waste disposal. The second downside aspect of polynuclear metal

complexes is that the fine tuning of the spectroscopic and chemical properties is usually harder

to achieve than for mononuclear catalysts, on which direct substitutions and changes in the

ligand environment make it easier to establish key structure-activity relationships and to study

mechanistic aspects of the catalysis: studies are greatly simplified if only one active metal site

is involved.

11

Ruthenium catalysts constitute the most studied family of catalysts for water oxidation, with

several mechanisms proposed to explain the observed behaviour.

Two of them, namely the water nucleophilic attack (WNA) mechanism and the radical coupling

(RC) mechanism, constitute the most widely accepted and supported according to experimental

evidence.

The water nucleophilic attack mechanism postulates that ligands with at least one labile

coordination site allow for the formation of a high valent [RuV=O]2+ species through several

PCET steps after initial coordination of a water molecule. This high valent species is then

attacked by the lone pair of the oxygen of another water molecule, forming a short lived

[RuIII-OOH]2+ intermediate, which loses one proton and one electron to form a [RuIV-OO]2+

intermediate, which then releases gaseous oxygen after electronic rearrangement and the

catalytic cycle is again started by coordination of a new water molecule. Key aspects of the water

nucleophilic attack mechanism have also been identified in iron, iridium and cobalt complexes14,

18, 19, as shown in Fig. 5.

Fig. 5: Summary of WNA-type mechanisms for single site catalysts based on ruthenium, iridium and cobalt.

Adapted from ref. 18 with permission.

12

The radical coupling mechanism (Fig. 6) postulates the formation via PCET of a high valent

oxyl radical, [RuIV-O•]2+, which is a resonance form of the [RuV=O]2+ species identified in the

WNA mechanism. This oxyl radical couples with another unit to yield a [RuIV-(O2)-RuIV] complex

that then releases dioxygen. The particular bonding mode of the oxygen and the last mechanistic

steps will depend on particular ligand geometry and environment, but the rate law has been

found to be second order with respect to the concentration of the catalyst, which strongly

supports the bimolecular nature of the mechanism.

Fig. 6: Radical coupling mechanism for single site catalysts based on ruthenium.

Adapted from ref. 18 with permission.

Previous catalysts based in pentadentate polypyridine ligands

Of the published systems, the one studied by Goldsmith et al.20, 21 and Wasylenko et al.22 was

chosen as a basis for this project. The original ligand framework consists of a five-coordinated

pentapyridine system: a central 2,6-disubstituted pyridine ring with four 2-pyridyl arms and

hydroxyl-terminated carbon linkers. Fig. 7 shows the crystal structure of PY5-OH,

pyridine-2,6-diylbis(di(pyridin-2-yl)methanol), closely related to the original PY5 ligand (with

MeOH groups). The crystal geometry does not differ significantly between both ligands.

Fig. 7: Crystal structure of PY5-OH ligand. Colour code: O (red), N (purple), C (grey), H (white).

Hydrogen bonds in light blue. X-Ray single-crystal diffraction data obtained from ref. 23 with permission.

13

When forming complexes with first row transition metals, the four pyridine arms bind to the

metal core with the four pyridine nitrogen atoms forming a plane, while the central metal-

pyridine bond lies perpendicular to this plane, leaving the sixth coordination position free for

the coordination of labile ligands (such as MeCN, Cl− and H2O). Furthermore, the metal-nitrogen

bond trans to this coordination site is greatly stabilized by means of the chelate effect22, making

the complex more stable and further improving the catalytic activity. The geometry is clearly

illustrated in the crystal structure of the complex [FeII(PY5-OH)Cl](CF3SO3) (Fig. 8):

Fig. 8: Two views of the crystal structure of [FeII(PY5-OH)Cl](CF3SO3). Colour code: Fe (orange), Cl (green), O (red),

N (purple), C (grey). Hydrogens and counter ions omitted for clarity, except for OH groups.

X-Ray single-crystal diffraction data obtained from ref. 24 with permission.

Synthetic routes for polypyridine ligands: considerations

Two main synthetic routes for PY5-type ligands have been proposed in the literature20-23, 25. Both

involve the addition of 2-lithiopyridine to carbonyl-containing pyridine moieties (i.e. dipicolinic

acid) in two to four steps. The advantages and disadvantages of each method are discussed

below.

1.7.1. Synthesis of pentapyridine ligands from 1,5-dipicolinic acid

The original synthesis of the PY5 ligand25 (Scheme 1) comprises the formation of the diacid

chloride derivative by reaction with SOCl2, subsequent reaction at low temperature with four

equivalents of 2-lithiopyridine (generated in-situ by lithium-halogen exchange at low

temperature) and further permethylation with methyl iodide. This synthetic route has been

reported to lead to several purification problems19, 20 and average overall yields of less than 50%.

Scheme 1: Original synthesis of PY5 ligand from dipicolinic acid

14

1.7.2. Synthesis of pentapyridine ligands from 1,5-dipicolinic acid diesters

A second synthetic route to the PY5 ligand (Scheme 2) was employed recently in the

literature20, 21 to improve the yield, facilitate the separation of the desired product and broaden

the scope of modifications to this ligand scaffold. In this synthetic route, dipicolinic acid alkyl

diesters are directly reacted with two equivalents of 2-lithiopyridine, from which a crystalline

solid intermediate diketone (PY3, see Scheme 2) is easily isolated by recrystallization in very

high yield (85%). This intermediate is further reacted with two more equivalents of 2-

lithiopyridine to yield the PY5-OH precursor, which is then permethylated to yield the PY5

ligand in greater overall yield (64%) while totally avoiding the need for purification via column

chromatography.

Scheme 2: Synthesis of PY5 ligand via PY3 diketone intermediate

The formation of the PY3 intermediate opens the ligand system for possible modifications, which

might consist, for example, on the introduction of two different “arms” for the ligand, such that

it would have two pyridine arms and two different arms (which might also be substituted

pyridines or other rings). The colour coding in Scheme 1 and Scheme 2 illustrates this

additional difference in the synthetic routes.

Proposed modifications to the ligand scaffold

The PY5 and PY5-OH ligands do not differ significantly in their geometry and metal binding

ability, and first row transition metal complexes from these ligands have been shown to be very

active as catalysts for water oxidation22, oxidation of unsaturated compounds20, 21, proton

reduction26 and even as cytotoxic anti-tumour agents24.

It has been shown22, 26 that the metal-ligand bond trans to the sixth coordination site (occupied

by water in the catalytic cycle) strongly affects the performance of the catalyst. If it is too weak,

then the catalyst might decompose into nanoparticles, while if it is too strong then the high-

valent oxo intermediate might not be electrophilic enough to promote the attack of water, or

might interfere with radical coupling. The trans influence can be accounted as the second major

source and explanation for this change in stability and reactivity.

15

The proposed modification to the ligand system consists in the replacement of the central

pyridine ring from PY5-OH for a five membered pyrrole ring to obtain the ligand shown in Fig.

9, (1H-pyrrole-2,5-diyl)bis(di(pyridin-2-yl)methanol) (PyPY4).

Fig. 9: Structure of the proposed ligand, (1H-pyrrole-2,5-diyl)bis(di(pyridin-2-yl)methanol)

This ligand is expected to introduce both a steric and electronic modification with respect to the

PY5 ligand. The NH group in pyrrole is to be deprotonated upon complexation, to form a N−(N)4

pentacoordinate ligand. The negative charge in the nitrogen from the pyrrole ring will then be

situated trans to the expected coordination site for water. Since the M+-N− has a more polar

character it is also expected to be stronger than the regular N→M dative bond, and is thus

expected to improve the stability of the complexes and increase the catalytic activity.

Synthetic approach to the proposed ligand

The two routes reported in literature for the preparation of PY5-OH and PY5 are, in principle,

viable. However, it has been reported27-31 that pyrrole-2-carboxylic acid readily undergoes

decarboxylation and is fairly unstable even under mildly acidic conditions, and therefore the

pyrrole-2,5-dicarboxylic acid and the diacid chloride derivative are also expected to be fairly

unstable. Furthermore, unprotected pyrrole is a substrate prone to polymerisation, which

suggests avoiding the diacid chloride route and the use of a protecting group for NH to prevent

polymerisation of the intermediate diketone and/or the starting pyrrole 2,5-diester. Boc was

chosen as the protecting group due to its easy introduction and removal32-35. Synthesis without

this protecting group was also attempted as the first choice. The proposed synthetic route for

the PyPY4 ligand is then a modification of the one reported for the PY5-OH ligand starting from

pyrrole-2,5-diester (Scheme 3).

Scheme 3: Proposed synthetic route for the new PyPY4 ligand

16

AIM OF THE PROJECT

The aim of this project was to synthesize a new ligand (PyPY4, (1H-pyrrole-2,5-

diyl)bis(di(pyridin-2-yl)methanol)), which would then be reacted with first row transition metal

salts to form new complexes that could show potential catalytic activity for water oxidation.

EXPERIMENTAL

Chemicals and solvents

Chemicals and solvents were obtained from commercial sources (Sigma-Aldrich) and were used

without further purification unless otherwise stated. Pyrrole was freshly distilled under vacuum

before use. Dry methanol, ethanol and CCl4 were obtained after storing the corresponding

solvent overnight under 15 wt.% freshly activated molecular sieves (4 Å). Dry THF and diethyl

ether were obtained by distillation under inert atmosphere from sodium/benzophenone still. All

glassware for water-sensitive reactions was stored overnight in the oven at 120 ºC.

Spectroscopic characterisation

NMR spectra were acquired on JEOL Eclipse+ ECP400 (1H 400 MHz, 13C 100 MHz) or Agilent

Mercury 400 (1H 400 MHz, 13C 100 MHz) spectrometers. All spectra were processed with

appropriate referencing to the solvent residual signal. For the measuring of small coupling

constants, where needed, the spectra were multiplied by the appropriate weighting functions

(exponential and gaussian) to resolve the peaks. DEPT135 and DEPT90 spectra (DEPT135

shows protonated CH/CH3 and CH2 carbons with opposite phases, and DEPT90 shows CH carbons

only, allowing for differentiation) were acquired for some of the compounds and used for the

assignment of the 13C resonances. LC-MS analysis were performed on a Dionex UltiMate 3000

LC system coupled to a Thermo Finnigan LCQ Deca XP Max ion trap mass spectrometer.

Synthesis of pyrrole-2,5-dicarboxylic acid diesters (dimethyl, diethyl)

In a dry microwave vial with a magnetic stir bar were added under argon Fe(acac)3 (3.2 mg, 0.01

mmol), pyrrole (60 mg, 0.9 mmol), dry CCl4 (1 g, 7 mmol) and dry MeOH (2 g, 60 mmol) or dry

EtOH (2.8 g, 60 mmol), which was then sealed hermetically with exclusion of oxygen by repeated

freeze-pump-thaw cycles, and reacted in a Biotage Initiator microwave reactor under different

reaction conditions, but the reaction yielded a black mixture of polypyrrole and less than 5% of

the desired product, among with an important (11%) fraction of the monoester.

17

Isolation of the reaction material from entry 9 in Table 1 by column chromatography (Silica;

0.01% Et3N in dichloromethane) yielded the monoester, 2-methyl-pyrrole-2-carboxylate (62 mg,

11%) as an off-white solid which readily decomposes upon storing for >1 week at RT. 1H NMR

(400 MHz, CDCl3, 25ºC) δ 9.77 (br. s, 1H; NH), 6.97 (td, J = 2.7, 1.5 Hz, 1H; H3 pyrrole), 6.94

(ddd, J = 3.7, 2.4, 1.5 Hz, 1H; H5 pyrrole), 6.26 (dt, J = 3.7, 2.5 Hz, 1H; H4 pyrrole), 3.86 (s, 3H;

OCH3). These results are in agreement with previously reported data for this compound36.

The reaction was also attempted with the same molar amount of CBr4 and some other catalysts,

and even without the use of microwave irradiation, yielding the same results (see discussion

below). This procedure was adapted from the reported by Khusnutdinov et al.37, where the

reaction yield is reported to be 99%.

Synthesis of Di(n-butyl) pyrrole-2,5-dicarboxylate

In an attempt to understand the reactivity of the system and get some insights into the

reactivity of pyrrole under the previous conditions, CBr4 (2.32 g, 7 mmol), pyrrole (60 mg, 0.9

mmol) and n-butanol (4.5 g, 60 mmol) were refluxed for 5 hours under inert atmosphere and

monitored by TLC and LC-MS, but only trace amounts of the expected di(n-butyl) pyrrole-2,5-

dicarboxylate were detected, while the reaction resulted in a black tar of polymeric material (see

discussion below).

Synthesis of N-Boc-pyrrole (tert-butyl pyrrole-1-carboxylate)

N-Boc pyrrole was synthesized as reported38. To a round bottomed flask containing pyrrole (2.0

g, 30 mmol) and 4-(dimethylamino)pyridine, DMAP, (500 mg, 4.5 mmol) in acetonitrile (30 mL)

was added di-tert-butyl dicarbonate, Boc2O, (7.8 g, 36 mmol) and the mixture stirred under argon

at room temperature for 2.5 hours. Evaporation of the solvent under reduced pressure and

purification by flash column chromatography (Al2O3; pentane) afforded 4.83 g (96%) of N-Boc-

pyrrole as a colourless liquid.

1H NMR (400 MHz, CDCl3, 25 ºC) δ 7.25 (t, J = 2.4 Hz, 2H; H2, H5), 6.22 (t, J = 2.4 Hz, 2H; H3,

H4), 1.61 (s, 9H; O-C(CH3)3).

Synthesis of 1-(tert-butyl) 2,5-dimethyl pyrrole-1,2,5-tricarboxylate

1-(tert-butyl) 2,5-dimethyl pyrrole-1,2,5-tricarboxylate was obtained from a modification of the

reported synthetic procedure39. In a 250 mL round-bottomed flask a 2.5 M solution of n-

butyllithium in hexanes (14.22 ml, 35.6 mmol) was slowly added to a solution of 2,2,6,6-

tetramethylpiperidine (6 ml, 35.6 mmol) in dry THF (50 mL) cooled to -78 °C under argon, to

give a light yellow solution, to which N-Boc-pyrrole (2.50 g, 15 mmol) in THF (10 ml) was added

18

dropwise and was left to stir for 3 hours at -78°C. The mixture was then transferred into a

dropping funnel and added dropwise into methyl chloroformate (3.5 ml, 45 mmol) and left to stir

at -78 ºC for an additional 30 minutes.

A saturated solution of ammonium chloride (10 mL) was then added and the mixture was

allowed to warm to room temperature. The mixture was then extracted with diethyl ether (25

mL) before being washed with 10% HCl solution (5 mL) and saturated sodium chloride solution

(20 mL).

The organic extracts were dried with MgSO4, filtered and concentrated under reduced pressure

to yield an orange solid, which was purified by flash column chromatography (Silica; 1:1

dichloromethane:pentane) and dried under vacuum overnight to yield pure

1-(tert-butyl) 2,5-dimethyl pyrrole-1,2,5-tricarboxylate as a white solid (3.56 g, 84%).

1H NMR (400 MHz, CDCl3, 25 ºC) δ 6.83 (s, 2H; H3, H4), 3.86 (s, 6H; 2xO-CH3), 1.66 (s, 9H; O-

C(CH3)3); 13C NMR (100 MHz, CDCl3) δ 160.02 (C=O(OMe)), 148.97 (C=O-O(t-Bu)), 126.82 (C2,

C5), 115.91 (C3, C4), 86.38 (O-C(CH3)3), 52.09 (O-CH3), 27.43 (t-Bu CH3).

Synthesis of methyl 5-(methoxydi(pyridin-2-yl)methyl)-pyrrole-2-

carboxylate (PyPY2)

To a 250 mL three-necked round-bottomed flask was added 2-bromopyridine (1.22 g, 7.8 mmol,

6 eq.) in dry THF (75 mL) to give a colourless solution, which was cooled in a methanol/liquid

nitrogen bath at -98 °C under argon. A 2.5 M solution of n-butyllithium in hexanes (3.10 mL,

7.8 mmol, 6 eq.) was added dropwise, keeping the reaction temperature below -80 ºC. A solution

of 1-(tert-butyl) 2,5-dimethyl pyrrole-1,2,5-tricarboxylate (365.4 mg, 1.3 mmol) in dry THF (20

mL) was added dropwise, keeping the temperature below -80 ºC. The reaction mixture was

stirred at -98 °C for 2 hours, after which the temperature was raised to RT and the reaction was

continuously stirred for 10 min. The reaction was quenched with MeOH (25 mL) and 5% aqueous

HCl (5 mL) was added, then the solution was basified with a saturated sodium hydroxide

solution (3 drops). The organic solvents were removed and the residue was extracted with

dichloromethane (5x10 mL), washed with brine and dried over Na2SO4. The organic extracts

were combined and dried under reduced pressure. An intensely colored deep red oil was then

purified by column chromatography (Al2O3; 0.1% Et3N in dichloromethane) to yield methyl 5-

(methoxydi(pyridin-2-yl)methyl)-pyrrole-2-carboxylate (PyPY2) as a deep red solid (243 mg,

38%). Further reaction of PyPY2 with two equivalents of 2-lithiopyridine did not afford the

desired PyPY4 product. See discussion below.

19

MS (ESI, m/z) 323.6 [M+] (Calc. for C18H17N3O3 323.1). 1H NMR (400 MHz, CDCl3, 25 ºC) δ 8.57

(ddd, J = 4.8, 1.8, 0.8 Hz, 2H; pyridine arms H6,H6’), 7.68 (ddd, J = 8.1, 7.5, 1.8 Hz, 2H; pyridine

arms H3,H3’), 7.58 (ddd, J = 8.1, 1.2, 0.8 Hz, 2H; pyridine arms H4,H4’), 7.16 (ddd, J = 7.5, 4.8,

1.2 Hz, 2H; pyridine arms H5,H5’), 6.88 (d, J = 3.9 Hz, 2H; pyrrole H3), 6.17 (d, J = 3.9 Hz, 2H;

pyrrole H4), 3.85 (s, 3H; OCH3 ester), 3.19 (s, 3H; OCH3). 13C NMR (100 MHz, CDCl3, 25 ºC) δ

162.47 (C=O ester), 161.71 (C2,C2’ pyridine arms), 148.65 (C6,C6’ pyridine arms), 137.57 (C5

pyrrole), 137.03 (C4,C4’ pyridine arms), 122.50 (C5,C5’ pyridine arms), 121.89 (C3,C3’ pyridine

arms), 121.71 (C2 pyrrole), 115.44 (C3 pyrrole), 110.91 (C4 pyrrole), 83.25 (C(OCH3)(PY)2),

52.90 (OCH3 ester), 51.40 (OCH3).

Synthesis of Nickel(II) and Copper(II) complexes with PyPY2

To a solution of PyPY2 (60.0 mg for Cu(II); 16.3 mg for Ni(II)) in methanol (0.5 mL) was added

under argon a solution of the corresponding metal bromide (CuBr2: 41.3 mg, 1 eq.; NiBr2 · 𝑥H2O:

13.4 mg, ~1 eq) in 0.5 mL MeOH. The reaction mixtures were stirred at room temperature for 4

hours after which LC-MS analysis were performed from the crude solutions, which turned deep

red for Ni(II) and dark green for Cu(II). No attempt was made to isolate the products. See

discussion below.

RESULTS AND DISCUSSION

Synthesis of pyrrole-2,5-dicarboxylic acid diesters (dimethyl, diethyl)

The attempted synthesis of pyrrole-2,5-dicarboxylic acid dimethyl and diethyl esters (Scheme

4) was not successful, even though special care was taken in drying the solvents and reagents

used for the reaction. Furthermore, the microwave vials were sealed and all reagents were

handled under argon atmosphere. Table 1 summarized the different reaction conditions that

were explored in an attempt to reproduce the results by Khusnutdinov et al.37 by using

microwave irradiation instead of thermal activation.

Scheme 4: Synthesis of pyrrole-2,5-dicarboxylic acid diesters

20

Trial Temp.

(ºC) Time

Pressure

(bar)

mmol

pyrrole

mmol

MeOH

mmol

𝐂𝐂𝐥𝟒

mmol

cat. Catalyst

Yield of

diester

1 115 5 min 10 0.9 60 7 0 none N.R.

2 120 5 min 10 0.9 60 7 0.01 Fe(acac)3 trace

3 150 5 min 18 0.9 60 7 0.01 Fe(acac)3 polymer

4 60 5 min 5 0.9 60 7 0.01 Fe(acac)3 N.R.

5 120 30 min 12 0.9 60 7 0.01 Fe(acac)3 trace

6 150 30 min 18 0.9 60 7 0.01 Fe(acac)3 polymer

7 120 1 h 18 4.5 350 35 0.05 Fe(acac)3 polymer

8 150 1 h 21 4.5 350 35 0.05 Fe(acac)3 polymer

9 120 30 min 12 4.5 350 35 0.05 FeCl2 <5%

10 150 30 min 18 4.5 350 35 0.05 FeCl2 trace

Table 1: Tested reaction conditions and yields for the synthesis of pyrrole-2,5-dicarboxylic acid dimethyl ester

under microwave irradiation. N.R = no reaction; polymer = polypyrrole was the main product as a black tar;

trace = only trace amounts were detected (by LC-MS).

This reaction was also attempted with EtOH and CBr4 under the same conditions as entries

1-5 in Table 1, but the desired product was detected only in trace amounts by LC-MS.

It is interesting to note that the reaction yield improved slightly by using FeCl2 as the catalyst

instead of Fe(acac)3. This fact can be explained due to the proven catalytic activity of Iron(III)

salts for the synthesis of polypyrrole40-43, and thus the use of Iron(II) salts might lower the yield

for the undesired polypyrrole, but is still inadequate because Iron(II) can be oxidized in solution

to Iron(III) probably by trace amounts of oxygen, thus generating the catalytic species that

promotes the undesired polymerisation.

The reaction mechanism for the esterification of pyrrole with CCl4 or CBr4 is thought to proceed

via radical addition to form a pyrrole-CX3 (X = Cl, Br) intermediate, which then undergoes

alcoholysis to afford the monoester. This product undergoes yet another radical addition and

alcoholysis in the 5 position of the ring, thus affording the diester. The mechanism is discussed

in more detail by Khusnutdinov et al.37

From these experiments, it is important to note that the formation of polymeric material, the

instability of the monoester, the difficulty in successful separation of the desired product, added

to the fact that the major product was the monoester instead of the diester, hindered further

attempts to synthesize the NH unprotected dimethyl pyrrole-2,5-dicarboxylate, and thus this

synthetic route was abandoned.

Synthesis of Di(n-butyl) pyrrole-2,5-dicarboxylate

A reaction with CBr4, n-butanol and pyrrole refluxed under inert atmosphere was carried out to

test conditions where the reagents were kept below their boiling point and without the use of

microwave irradiation (Scheme 5).

21

Scheme 5: Synthesis of di(n-butyl) pyrrole-2,5-dicarboxylate

However, the formation of the expected di(n-butyl) pyrrole-2,5-dicarboxylate was achieved only

in trace quantities, while the reaction mixture turned deep black after 2.5 hours of refluxing,

indicating the presence of a big amount of polypyrrole. TLC and LC-MS analysis indicated the

formation of other side products besides polypyrrole, which constitute a final confirmation that

raw pyrrole is not a good substrate to start with in order to get 2,5-diesters.

Synthesis of N-Boc-pyrrole

In this reaction (Scheme 6), the excellent yield, stability and ease of purification of the product

contrast with the behaviour of deprotected pyrrole, thus suggesting that the protection of the

NH group of pyrrole was, in fact, needed for this synthetapproach to succeed. Boc2O is a cheap

reagent and the reaction conditions are straightforward. Note however that N-Boc-pyrrole is

also available commercially.

Scheme 6: Synthesis of N-Boc-pyrrole

Synthesis of 1-(tert-butyl) 2,5-dimethyl pyrrole-1,2,5-tricarboxylate

The modification of the reported synthesis introduced in this work consisted of using a dropping

funnel instead of a cannula to transfer the lithiated pyrrole intermediate into methyl

chloroformate. This reaction is a model reaction for the attack of N-Boc 2,5-dilithiopyrrole to an

electrophilic substrate, methyl chloroformate, according to Scheme 7.

Scheme 7: Synthesis of 1-(tert-butyl) 2,5-dimethyl pyrrole-1,2,5-tricarboxylate

An important aspect of this reaction is that the N-Boc 2,5-dilithiopyrrole intermediate is

stabilised by the action of the Boc protecting group, which has an impact on the stability of this

organolithium reagent, as illustrated in Fig. 10.

22

Fig. 10: Stabilisation of 2,5-dilithiopyrrole by the N-Boc protecting group

Synthetic routes to the formation of N-Boc 2,5-dilithiopyrrole also include the lithium-halogen

exchange, but in this particular case, the reaction of n-BuLi with HTMP (2,2,6,6-

tetramethylpiperidine) to afford LiTMP (lithium 2,2,6,6-tetramethylpiperidide) and subsequent

reaction with N-Boc pyrrole enable for the synthesis of the 2,5-dimethyl ester. A final remark

from this reaction is that the Boc protecting group was not cleaved or attacked under these

reaction conditions.

Synthesis of methyl 5-(methoxydi(pyridin-2-yl)methyl)-pyrrole-2-

carboxylate (PyPY2)

Initial attempts to synthesize an intermediate analogous to PY3 (Scheme 2) by using only two

equivalents of 2-lithiopyridine in dry THF at -78 ºC were unsuccessful, and only the starting

triester was isolated from the reaction mixture.

The reaction was also attempted in dry Et2O at -78 ºC, but this solvent promotes decomposition

of the 2-lithiopyridine reagent, and for that particular case, no traces of the starting material

could be isolated, polymeric material being the only product. When four equivalents of 2-

lithiopyridine were used, the reaction yield of PyPY2 improved slightly to ~15%, but still the

desired PyPY4 product was not observed.

To improve the stability of the 2-lithiopyridine, very dilute solutions and much lower

temperature (-98 ºC) were tested, which improved the reaction yield and allowed the successful

synthesis of PyPY2, according to Scheme 8.

Scheme 8: Synthesis of PyPY2

23

A remarkable feature of this reaction is that the obtained product is an asymmetrically

substituted pyrrole, which contrasts with the symmetric pyridine diketone intermediate (PY3)

reported when pyridine 2,6-diesters are used as the starting materials. This suggests that

pyrrole 2,5-diesters have a very different reactivity trend towards addition of 2-lithiopyridine.

Also, methoxy groups were found in the final product instead of the expected OH groups.

Furthermore, when six equivalents of 2-lithiopyridine were used, PyPY2 was the only product

observed and the reaction yield improved significantly (38%). Only polymeric material was

observed when PyPY2 was reacted again with two equivalents of 2-lithiopyridine, which

confirms that six equivalents is a fair excess.

The 13C chemical shift in CDCl3 of the methoxy ester carbonyl in 1-(tert-butyl)-2,5-dimethyl

pyrrole-1,2,5-tricarboxylate is 160.08 ppm, while that of the pyridine analogue (dimethyl

dipicolinate) is 164.6 ppm44, thus suggesting that the two carbonyls do not differ significantly

in electrophilicity, while still it can be deduced that the carbonyl from the pyrrole system is

more shielded (and therefore less electrophilic) because the pyrrole ring is more electron-rich

than the pyridine ring. A final interesting aspect of this reaction is that the Boc protecting group

is removed from the pyrrole system during workup, as determined by LC-MS, even though the

NH resonance was not found in the 1H NMR spectrum due to the sample being too dilute (thus

making the signal-to-noise ratio for the broad NH resonance too low to be detected). We

therefore propose that the Boc protecting group is too bulky to allow attack of the 2-

lithiopyridine on the second carbonyl once the first pyridine has been introduced.

Attempted synthesis of Nickel(II) and Copper(II) complexes of the

PyPY2 ligand

The fact that PyPY2 is a new tridentate ligand itself prompted for an attempt to synthesize

complexes with first row transition metals. Among them Nickel(II) and Copper(II) were chosen

since they are known to form stable tetracoordinated complexes.

The attempted synthesis of a Copper(II) complex yielded a mixture of complexes, identified by

mass spectrometry, but whose structural characteristics such as the disposition of the ligands

around the metal centre need to be further investigated, and to improve the reaction yield the

mixture should be refluxed and a catalytic amount of a weak base added to promote

deprotonation of the pyrrole NH group. Fig. 11 illustrates the experimental and calculated mass

spectrum of a Cu(II) complex with candidate formula [Cu(PyPY2−)2], resulting from the

coordination of two ligand molecules to a single metal site. A candidate structure for this

complex is shown in Fig. 12, but further spectroscopic studies (XRD, NMR) are needed in order

to confirm the structures and oxidation states.

24

Fig. 11: Mass spectrum of the complex ion with formula [Cu(PyPY2−)2], C36H33CuN6O6.

Top: experimental mass spectrum; Bottom: calculated isotopic distribution for C36H34CuN6O6+, [M+H]+.

Fig. 12: Candidate structure for complex [Cu(PyPY2−)2] based on mass spectrometric results

Also, another mass pattern was recognized (Fig. 13), corresponding to an oxo bridged Cu(III)

dimer, [Cu(PyPY2−)(μ-O)2(PyPY2−)Cu]. The candidate structure is shown in Fig. 14.

Fig. 13: Mass spectrum of the complex ion with formula [Cu(PyPY2−)(μ-O)2(PyPY2−)Cu], C36H32Cu2N6O8.

Top: experimental mass spectrum; Bottom: calculated isotopic distribution for C36H33Cu2N6O8+, [M+H]+.

700 705 710 715 720

m/z

0

20

40

60

80

100

0

20

40

60

80

100

Re

lative

Ab

un

da

nce

708.99

710.94

712.02

712.93704.25700.85 715.30 717.61706.59

709.18

711.18

712.18

713.19 715.19 718.20

NL:1.75E8

RF24A-Cu-after4h#211 RT: 5.64 AV: 1 T: + c ESI Full ms [80.00-2000.00]

NL:4.51E5

C 36 H 34 CuN 6 O 6: C 36 H 34 Cu 1 N 6 O 6

pa Chrg 1

790 795 800 805 810

m/z

0

20

40

60

80

100

0

20

40

60

80

100

Rela

tive A

bundance

802.92

804.98

806.01

808.19

796.24 801.92 813.70811.53799.62787.97 790.48

803.09805.09

806.10

807.09

808.09810.10 813.11

NL:3.80E7

RF24A-Cu-after4h#255 RT: 6.81 AV: 1 T: + c ESI Full ms [80.00-2000.00]

NL:3.10E5

C 36 H 33 Cu 2 N 6 O 8+: C 36 H 33 Cu 2 N 6 O 8

pa Chrg 1

25

Fig. 14: Candidate structure for complex [Cu(PyPY2−)(μ-O)2(PyPY2−)Cu]+ based on mass spectrometric results

The high valent Cu(III) structure is suggested from oxo bridges, which are favoured for higher

oxidation states.

The attempted synthesis of a Nickel complex yielded new species in the mass chromatogram,

but no recognizable mass patterns were detected. As for the case of Copper, the reaction mixture

should be refluxed and a catalytic amount of a weak base added to promote deprotonation of the

pyrrole NH group, and after obtaining the corresponding complexes, further spectral

characterisation is needed.

These complexes were not tested for water oxidation, but Ni(II) and Cu(II) complexes have also

been shown very recently to possess catalytic activity towards water oxidation.

CONCLUSIONS

In conclusion, the synthesis of a new pyrrole-based ligand, methyl 5-(methoxydi(pyridin-2-

yl)methyl)-pyrrole-2-carboxylate (PyPY2), was achieved under moderate yield (38%) by addition

of six equivalents of 2-lithiopyridine at -98ºC in THF to 1-(tert-butyl) 2,5-dimethyl pyrrole-1,2,5-

tricarboxylate. Deprotection of the pyrrole ring was observed during workup, thus making a

deprotection step unnecessary. The ligand was properly characterized by 1H and 13C NMR

spectroscopy and mass spectrometry. Attempted synthesis of pyrrole 2,5-dicarboxylic acid

diesters from reported methods were not successful under the tested conditions, and thus the

Boc protecting group was shown to be necessary for the success of the reaction. Attempted

synthesis of the original proposed ligand, (1H-pyrrole-2,5-diyl)bis(di(pyridin-2-yl)methanol)),

was not successful. The obtained PyPY2 compound could instead be used as a tridentate ligand,

tentatively forming coordination compounds with Ni(II) and Cu(II), which were identified by

LC-MS in the case of Cu(II). Further characterization and analysis of the ligand and its metal

complexes are a work still on progress, and in the future these complexes should be tested for

their catalytic activity towards water oxidation.

26

ACKNOWLEDGEMENTS

I would like to thank Dr. Anders Thapper and Dr. Sascha Ott for accepting me in their research

group, thus allowing me to carry out this project and enabling me to learn more about transition

metal chemistry and water oxidation catalysis in the labs. I would also like to thank Dr.

Biswanath Das for his great help in developing this project, ranging from the introduction to

the PY5-OH system to key discussions, help and encouragement throughout the whole project.

Many thanks have to be given as well to all the administrative and support personnel in House

7 at Ångström laboratory.

I wish to thank my family, especially my mother, whose unconditional support, even throughout

the distance, helped me to be strong even in the toughest moments.

To my best friend, Estefanía Sucre, and the rest of my friends from the Laboratory for

Photochemistry and Photobiology at Universidad Simón Bolívar in Venezuela: María Lourdes

Nóvoa, Carlos Gámez, Prof. Nieves Canudas and Prof. Sara Pekerar for their support and

encouragement, many thanks.

To my professors from the Department of Chemistry at Universidad Simón Bolívar in

Sartenejas, Venezuela, who formed me during four years and enabled me to lift my wings for

this dream which is now a reality. Thank you all.

Finally, I wish to thank the great academic composers: Sergei Rachmaninoff, Franz Liszt, Pyotr

Ilyich Tchaikovsky, Frédéric Chopin, Wolfgang Amadeus Mozart, Charles Camille Saint-Saëns

and many others, whose wonderful music was essential for the writing of this manuscript and

has accompanied me during almost my whole life.

“Success is not final, failure is not fatal: it is the courage to continue that counts.”

Winston Churchill

27

REFERENCES

1. Panwar, N.L., S.C. Kaushik, and S. Kothari, Role of renewable energy sources in

environmental protection: A review. Renewable and Sustainable Energy Reviews, 2011.

15(3): p. 1513-1524, DOI: 10.1016/j.rser.2010.11.037.

2. Pérez-Lombard, L., J. Ortiz, and C. Pout, A review on buildings energy consumption

information. Energy and Buildings, 2008. 40(3): p. 394-398, DOI:

10.1016/j.enbuild.2007.03.007.

3. Solomon, S., G.K. Plattner, R. Knutti, and P. Friedlingstein, Irreversible climate change

due to carbon dioxide emissions. Proc Natl Acad Sci U S A, 2009. 106(6): p. 1704-9, DOI:

10.1073/pnas.0812721106.

4. Kasting, J.F. and J.L. Siefert, Life and the evolution of Earth's atmosphere. Science, 2002.

296(5570): p. 1066-8, DOI: 10.1126/science.1071184.

5. Collings, A.F. and C. Critchley, Artificial photosynthesis: from basic biology to industrial

application. 2005, Weinheim: Wiley-VCH.

6. Kärkäs, M.D., O. Verho, E.V. Johnston, and B. Åkermark, Artificial photosynthesis:

molecular systems for catalytic water oxidation. Chem Rev, 2014. 114(24): p. 11863-2001,

DOI: 10.1021/cr400572f.

7. Najafpour, M.M., A.N. Moghaddam, S.I. Allakhverdiev, and Govindjee, Biological water

oxidation: lessons from nature. Biochim Biophys Acta, 2012. 1817(8): p. 1110-21, DOI:

10.1016/j.bbabio.2012.04.002.

8. Yano, J., J. Kern, K. Sauer, M.J. Latimer, Y. Pushkar, J. Biesiadka, B. Loll, W. Saenger,

J. Messinger, A. Zouni, and V.K. Yachandra, Where water is oxidized to dioxygen:

structure of the photosynthetic Mn4Ca cluster. Science, 2006. 314(5800): p. 821-5, DOI:

10.1126/science.1128186.

9. Ferreira, K.N., T.M. Iverson, K. Maghlaoui, J. Barber, and S. Iwata, Architecture of the

photosynthetic oxygen-evolving center. Science, 2004. 303(5665): p. 1831-8, DOI:

10.1126/science.1093087.

10. Yano, J. and V. Yachandra, Mn4Ca cluster in photosynthesis: where and how water is

oxidized to dioxygen. Chem Rev, 2014. 114(8): p. 4175-205, DOI: 10.1021/cr4004874.

11. Tran, R., J. Kern, J. Hattne, S. Koroidov, J. Hellmich, R. Alonso-Mori, N.K. Sauter, U.

Bergmann, J. Messinger, A. Zouni, J. Yano, and V.K. Yachandra, The Mn4Ca

photosynthetic water-oxidation catalyst studied by simultaneous X-ray spectroscopy and

crystallography using an X-ray free-electron laser. Philos Trans R Soc Lond B Biol Sci,

2014. 369(1647): p. 20130324, DOI: 10.1098/rstb.2013.0324.

12. Zhang, C., C. Chen, H. Dong, J.R. Shen, H. Dau, and J. Zhao, A synthetic Mn4Ca-cluster

mimicking the oxygen-evolving center of photosynthesis. Science, 2015. 348(6235): p. 690-

3, DOI: 10.1126/science.aaa6550.

13. Sun, L., A closer mimic of the oxygen evolution complex of photosystem II. Science, 2015.

348(6235): p. 635-6, DOI: 10.1126/science.aaa9094.

28

14. Limburg, B., E. Bouwman, and S. Bonnet, Molecular water oxidation catalysts based on

transition metals and their decomposition pathways. Coordination Chemistry Reviews,

2012. 256(15-16): p. 1451-1467, DOI: 10.1016/j.ccr.2012.02.021.

15. Cady, C.W., R.H. Crabtree, and G.W. Brudvig, Functional Models for the Oxygen-

Evolving Complex of Photosystem II. Coord Chem Rev, 2008. 252(3-4): p. 444-455, DOI:

10.1016/j.ccr.2007.06.002.

16. Gersten, S.W., G.J. Samuels, and T.J. Meyer, Catalytic oxidation of water by an oxo-

bridged ruthenium dimer. Journal of the American Chemical Society, 1982. 104(14): p.

4029-4030, DOI: 10.1021/ja00378a053.

17. Duan, L., F. Bozoglian, S. Mandal, B. Stewart, T. Privalov, A. Llobet, and L. Sun, A

molecular ruthenium catalyst with water-oxidation activity comparable to that of

photosystem II. Nat Chem, 2012. 4(5): p. 418-23, DOI: 10.1038/nchem.1301.

18. Wasylenko, D.J., R.D. Palmer, and C.P. Berlinguette, Homogeneous water oxidation

catalysts containing a single metal site. Chem Commun (Camb), 2013. 49(3): p. 218-27,

DOI: 10.1039/c2cc35632e.

19. Kärkäs, M.D., T. Åkermark, E.V. Johnston, S.R. Karim, T.M. Laine, B.L. Lee, T.

Åkermark, T. Privalov, and B. Åkermark, Water oxidation by single-site ruthenium

complexes: using ligands as redox and proton transfer mediators. Angew Chem Int Ed

Engl, 2012. 51(46): p. 11589-93, DOI: 10.1002/anie.201205018.

20. Goldsmith, C.R. and T.D. Stack, Hydrogen atom abstraction by a mononuclear ferric

hydroxide complex: insights into the reactivity of lipoxygenase. Inorg Chem, 2006. 45(15):

p. 6048-55, DOI: 10.1021/ic060621e.

21. Goldsmith, C.R., R.T. Jonas, and T.D.P. Stack, C−H Bond Activation by a Ferric

Methoxide Complex: Modeling the Rate-Determining Step in the Mechanism of

Lipoxygenase. Journal of the American Chemical Society, 2002. 124(1): p. 83-96, DOI:

10.1021/ja016451g.

22. Wasylenko, D.J., C. Ganesamoorthy, J. Borau-Garcia, and C.P. Berlinguette,

Electrochemical evidence for catalytic water oxidation mediated by a high-valent cobalt

complex. Chem Commun (Camb), 2011. 47(14): p. 4249-51, DOI: 10.1039/c0cc05522k.

23. Görlitz, J.J., P. Nielsen, H. Toftlund, and A.D. Bond, 2,6-Bis[bis(2-

pyridyl)hydroxymethyl]pyridine. Acta Crystallographica Section E Structure Reports

Online, 2004. 60(8): p. o1319-o1320, DOI: 10.1107/s1600536804016368.

24. Wong, E.L., G.S. Fang, C.M. Che, and N. Zhu, Highly cytotoxic iron(II) complexes with

pentadentate pyridyl ligands as a new class of anti-tumor agents. Chem Commun (Camb),

2005(36): p. 4578-80, DOI: 10.1039/b507687k.

25. Jonas, R.T. and T.D.P. Stack, C−H Bond Activation by a Ferric Methoxide Complex: A

Model for the Rate-Determining Step in the Mechanism of Lipoxygenase. Journal of the

American Chemical Society, 1997. 119(36): p. 8566-8567, DOI: 10.1021/ja971503g.

26. Bachmann, C., M. Guttentag, B. Spingler, and R. Alberto, 3d element complexes of

pentadentate bipyridine-pyridine-based ligand scaffolds: structures and photocatalytic

activities. Inorg Chem, 2013. 52(10): p. 6055-61, DOI: 10.1021/ic4004017.

27. Alvarez-Builla, J., J.J. Vaquero, and J. Barluenga, Modern heterocyclic chemistry. 2011,

Weinheim: Wiley-VCH.

29

28. Dunn, G.E. and G.K.J. Lee, Kinetics and Mechanism of the Decarboxylation of Pyrrole-2-

carboxylic Acid in Aqueous Solution. Canadian Journal of Chemistry, 1971. 49(7): p.

1032-1035, DOI: 10.1139/v71-172.

29. Liang, J., B. Wang, and Z. Cao, The Mechanism of Acid-Catalyzed Decarboxylation of

Pyrrole-2-Carboxylic Acid: Insights from Cluster-Continuum Model Calculations.

Journal of Theoretical and Computational Chemistry, 2013. 12(04): p. 1350017, DOI:

10.1142/s021963361350017x.

30. Cheng, X., J. Wang, K. Tang, Y. Liu, and C. Liu, Decarboxylation of pyrrole-2-carboxylic

acid: A DFT investigation. Chemical Physics Letters, 2010. 496(1-3): p. 36-41, DOI:

10.1016/j.cplett.2010.07.052.

31. Mundle, S.O. and R. Kluger, Decarboxylation via addition of water to a carboxyl group:

acid catalysis of pyrrole-2-carboxylic acid. J Am Chem Soc, 2009. 131(33): p. 11674-5,

DOI: 10.1021/ja905196n.

32. Wang, J., Y.L. Liang, and J. Qu, Boiling water-catalyzed neutral and selective N-Boc

deprotection. Chem Commun (Camb), 2009(34): p. 5144-6, DOI: 10.1039/b910239f.

33. Gibson, F.S., S.C. Bergmeier, and H. Rapoport, Selective Removal of an N-BOC

Protecting Group in the Presence of a tert-Butyl Ester and Other Acid-Sensitive Groups.

The Journal of Organic Chemistry, 1994. 59(11): p. 3216-3218, DOI:

10.1021/jo00090a045.

34. Han, G., M. Tamaki, and V.J. Hruby, Fast, efficient and selective deprotection of the tert-

butoxycarbonyl (Boc) group using HCl/dioxane (4 M). Journal of Peptide Research, 2001.

58(4): p. 338-341, DOI: 10.1034/j.1399-3011.2001.00935.x.

35. Ravinder, K., A.V. Reddy, K.C. Mahesh, M. Narasimhulu, and Y. Venkateswarlu, Simple

and Selective Removal of t-Butyloxycarbonyl (Boc) Protecting Group on Indoles, Pyrroles,

Indazoles, and Carbolines. Synthetic Communications, 2007. 37(2): p. 281-287, DOI:

10.1080/00397910601033724.

36. Schmuck, C., V. Bickert, M. Merschky, L. Geiger, D. Rupprecht, J. Dudaczek, P. Wich,

T. Rehm, and U. Machon, A Facile and Efficient Multi-Gram Synthesis of N-Protected 5-

(Guanidinocarbonyl)-1H-pyrrole-2-carboxylic Acids. European Journal of Organic

Chemistry, 2008. 2008(2): p. 324-329, DOI: 10.1002/ejoc.200700756.

37. Khusnutdinov, R.I., A.R. Baiguzina, R.R. Mukminov, I.V. Akhmetov, I.M. Gubaidullin,

S.I. Spivak, and U.M. Dzhemilev, New synthesis of pyrrole-2-carboxylic and pyrrole-2,5-

dicarboxylic acid esters in the presence of iron-containing catalysts. Russian Journal of

Organic Chemistry, 2010. 46(7): p. 1053-1059, DOI: 10.1134/s1070428010070158.

38. Salman, H., Y. Abraham, S. Tal, S. Meltzman, M. Kapon, N. Tessler, S. Speiser, and Y.

Eichen, 1,3-Di(2-pyrrolyl)azulene: An Efficient Luminescent Probe for Fluoride.

European Journal of Organic Chemistry, 2005. 2005(11): p. 2207-2212, DOI:

10.1002/ejoc.200500012.

39. Donohoe, T.J., C.E. Headley, R.P. Cousins, and A. Cowley, Flexibility in the partial

reduction of 2,5-disubstituted pyrroles: application to the synthesis of DMDP. Org Lett,

2003. 5(7): p. 999-1002, DOI: 10.1021/ol027504h.

40. Armes, S.P., Optimum reaction conditions for the polymerization of pyrrole by iron(III)

chloride in aqueous solution. Synthetic Metals, 1987. 20(3): p. 365-371, DOI:

10.1016/0379-6779(87)90833-2.

30

41. Kang, E.T., K.G. Neoh, Y.K. Ong, K.L. Tan, and B.T.G. Tan, X-ray photoelectron

spectroscopic studies of polypyrrole synthesized with oxidative iron(III) salts.

Macromolecules, 1991. 24(10): p. 2822-2828, DOI: 10.1021/ma00010a028.

42. Chitte, H.K., G.N. Shinde, N.V. Bhat, and V.E. Walunj, Synthesis of Polypyrrole Using

Ferric Chloride (FeCl&lt;sub&gt;3&lt;/sub&gt;) as Oxidant Together with Some Dopants

for Use in Gas Sensors. Journal of Sensor Technology, 2011. 01(02): p. 47-56, DOI:

10.4236/jst.2011.12007.

43. Vernitskaya, T.y.V. and O.N. Efimov, Polypyrrole: a conducting polymer; its synthesis,

properties and applications. Russian Chemical Reviews, 1997. 66(5): p. 443-457, DOI:

10.1070/RC1997v066n05ABEH000261.

44. Jokela, R., Interpretation of the 13C NMR spectra of thirteen methoxycarbonylpyridines

and some of their derivatives. Magnetic Resonance in Chemistry, 1992. 30(7): p. 681-683,

DOI: 10.1002/mrc.1260300720.

31

APPENDIXES

1H NMR spectrum of N-Boc pyrrole

1H, 13C and DEPT135 NMR spectra of 1-(tert-butyl) 2,5-dimethyl

pyrrole-1,2,5-tricarboxylate

1H, 13C, DEPT135 and DEPT90 NMR spectra of PyPY2

1H NMR spectrum of methyl 1H-pyrrole-2-carboxylate

32

8.1

33

8.2

34

8.2

35

8.3

36

8.3

37

8.3

38

8.3

39

8.4